Abstract
Multicellular organisms rely upon diverse and complex intercellular communications networks for a myriad of physiological processes. Disruption of these processes is implicated in the onset and propagation of disease and disorder, including the mechanisms of senescence at both cellular and organismal levels. In recent years, secreted extracellular vesicles (EVs) have been identified as a particularly novel vector by which cell-to-cell communications are enacted. EVs actively and specifically traffic bioactive proteins, nucleic acids, and metabolites between cells at local and systemic levels, modulating cellular responses in a bidirectional manner under both homeostatic and pathological conditions. EVs are being implicated not only in the generic aging process, but also as vehicles of pathology in a number of age-related diseases, including cancer and neurodegenerative and disease. Thus, circulating EVs—or specific EV cargoes—are being utilised as putative biomarkers of disease. On the other hand, EVs, as targeted intercellular shuttles of multipotent bioactive payloads, have demonstrated promising therapeutic properties, which can potentially be modulated and enhanced through cellular engineering. Furthermore, there is considerable interest in employing nanomedicinal approaches to mimic the putative therapeutic properties of EVs by employing synthetic analogues for targeted drug delivery. Herein we describe what is known about the origin and nature of EVs and subsequently review their putative roles in biology and medicine (including the use of synthetic EV analogues), with a particular focus on their role in aging and age-related brain diseases.
Keywords: Extracellular vesicles, Exosomes, Neurodegeneration, Aging, Drug delivery, Nanomedicine
Introduction
Safe, efficacious and specific drug delivery is integral to modern therapeutic medicine. The ability to optimise the bioavailability, stability, and targeted uptake of a therapeutic agent while simultaneously mitigating toxicity, immunogenicity and off-target/side effects is of utmost priority to in development of more effective drugs, and in the treatment of otherwise incurable diseases. Extensive efforts are being made in the modification or derivation of existing drugs, or the development of new drug-delivery platforms, to achieve these goals, often inspired by physiological mechanisms.
One such phenomenon is that of extracellular vesicles (EVs), a type of naturally occurring nanovesicles that envelop, protect and shuttle their bioactive cargo between cells in different systems (Thery 2011). These extracellular organelles are no longer considered to be mere cellular debris (Cocucci et al. 2009), nor are they just being proposed as circulating diagnostic markers that mirror their parental cell’s physiologic statuses, rather they appear to be central players in a diverse, complex, and specific intercellular communication network (Simons and Raposo 2009). As EVs are implicated in a plethora of physiological and pathological processes, a thorough understanding their origin and function is of great importance to medical science. Furthermore, their role as natural molecular cargo carriers provides inspiration for the design of new and improved therapeutic platforms, be they emulating EVs or repurposing them for medicinal applications.
Herein we review the current state of knowledge of EVs, describing their various classes, and providing examples of their function in disease, health, and during the processes of brain aging. A broad overview of the therapeutic potential of EVs is also provided, as is a rundown of current synthetic nanotherapeutic drug-delivery platforms that mimic the properties of EVs. While the field of EV study is still largely in its infancy, the therapeutic potential of EVs (and their analogues) in aging and age-related disease, particularly neurodegeneration, is plain to see.
Extracellular vesicles (EVs)
Characterisation of EVs
EV is a broad term used to describe membrane structures secreted by cells into the extracellular space to be later taken up by an target/acceptor cell (Raposo and Stoorvogel 2013). Despite the lack of definitive evidence for their physiological function in vivo, EVs appear to constitute a newly recognized means of communication found to be shared by almost every cell type (Thery 2011).
While the description of EVs has historically been burdened by a Byzantine nomenclature (Gould and Raposo 2013), a systematic classification based on the mechanisms of biogenesis and release of EVs (Akers et al. 2013) allows for the categorization of EVs into four broad groups:
-
(i)
Exosomes homogenous saucer-shaped EVs 30–100 nm in diameter, from multivesicular bodies (MVBs) of the endosomal pathway;
-
(ii)
Shedding vesicles (or microvesicles) heterogeneous EVs 50–2,000 nm in diameter, from direct blebbing of the cellular plasma membrane;
-
(iii)
Retrovirus-like particles (RLPs) sized 90–100 nm, with a typical subset of retroviral proteins but non-infectious, due to the lack of genes required for full viral propagation; and
-
(iv)
Apoptotic bodies 50–5,000 nm in diameter, vesicles arising during the apoptotic fragmentation of cells.
Other classes of EVs that fall outside these classifications have recently been identified. For instance, gesicles, approximately 100 nm in diameter and slightly less dense than exosomes, are highly fusogenic due to their origins in cells induced to overexpress the spike glycoprotein of the vesicular stomatitis virus (VSV-G) (Mangeot et al. 2011). Moreover, exosome-like vesicles (20–50 nm) expressing the full-length 55-kDa tumour necrosis factor (TNF) receptor 1 have been identified and may originate from multivesicular internal compartments (not necessarily being part of the endosomal system), but their nature is not well defined (Hawari et al. 2004).
Considering that a single cell type can secrete multiple EV classes (Heijnen et al. 1999; Deregibus et al. 2007; Muralidharan-Chari et al. 2009), one of the key challenges in the field is to establish methods allowing for their discrimination and—in perspective—their characterization and fractionation. Differences in properties such as size, morphology and density are not fully sufficient for a clear distinction (Bobrie et al. 2011). Further characterization requires biochemistry, qualitative and quantitative protein, RNA and lipid characterization, and imaging such as electron microscopy. Complementary to that, nano-particle-tracking analysis allows for the determination of EV size distribution based on the Brownian motion of vesicles in suspension (Soo et al. 2012). Furthermore, a novel high-resolution flow cytometry–based approach has been developed for quantitative high throughput analysis of immunolabeled vesicles (Nolte-’t Hoen et al. 2012; van der Vlist et al. 2012).
Nevertheless, while there is promiscuity in the expression of protein markers between EV classes, distinct combinations of markers are used to distinguish between different types of EVs. Exosomes are characteristically enriched into endosome-associated proteins [e.g., Rab GTPase, Soluble NSF Attachment Protein (SNAP) receptors (SNAREs), annexins, and flotillin], some of which are involved in MVB biogenesis (e.g., Alix and Tsg101 van Niel et al. 2006). CD63 and CD9, members of the tetraspanin family (Hemler 2003), are also potential markers of exosomes (Escola et al. 1998; Bard et al. 2004). Moreover, compared with plasma membrane-derived vesicles, exosomes are highly enriched in cholesterol, sphingomyelin, and hexosylceramides, at the expense of phosphatidylcholine and phosphatidylethanolamine (Wubbolts et al. 2003; Laulagnier et al. 2004; Subra et al. 2007; Brouwers et al. 2013). Furthermore, the constituent fatty acids of exosomes are primarily saturated or monounsaturated.
General markers of microvesicles are less well-defined, perhaps due to the diversity inherent in this class, but recently ADP-ribosylation factor 6 (ARF6) and vesicle-associated membrane protein 3 (VAMP3) have been proposed as potential candidates (Muralidharan-Chari et al. 2009). Shedding vesicles, ostensibly a sub-type of microvesicle, typically exhibit high levels of phosphatidylserine and are enriched in lipid raft-associated proteins such as tissue factor and flotillin-1, as well as various selectins and integrins, CD40 ligand, complement receptor-1, and the matrix metalloproteinases (MMP)-2 and -9 (Lee et al. 2011; Théry et al. 2009).
Retrovirus-like particles are less well studied, though Gag protein (together with other endogenous viral proteins) may be a general marker (Bronson et al. 1979; Boller et al. 1993; Mueller-Lantzsch et al. 1993; Dewannieux et al. 2005).
Finally, thrombospondin (TSP), complement subunit C3b and annexin V (all bound by phagocytes for the final clearance), together with histones and fragments of genomic DNA, are generally accepted markers of apoptotic bodies (Théry et al. 2009).
Biogenesis of EVs
Exosomes are formed in MVBs (El-Andaloussi et al. 2013), whereas microvesicles originate by direct budding from the plasma membrane (Raposo and Stoorvogel 2013). Thus, the overall molecular machineries involved in their formation and release are likely to be different (Fig. 1). Nevertheless, it should be noted that some aspects of their biogenesis might overlap. For instance, it has been suggested that microvesicle generation may necessitate factors also involved in exosome generation (Nabhan et al. 2012). Specifically, it was observed that a class of microparticles known as arrestin domain-containing protein 1 (ARRDC1)-mediated microvesicles (ARMMs) form and bud from the plasma membrane following an interaction between the tumour susceptibility gene 101 (TSG101), an endosome-associated protein implicated in exosome formation, and ARRDC1, localised to the plasma membrane. Moreover, actin-myosin interactions seem to play a critical role in the formation of all four types of EVs described above (Gladnikoff et al. 2009; Piper and Katzmann 2007; Sebbagh et al. 2001; Muralidharan-Chari et al. 2009). For instance, increased phosphorylation of the myosin light chain (MLC) has been shown to promote the actin-myosin contraction force leading to membrane blebbing; inhibitors of the MLC kinase were found to decrease blebbing (Mills et al. 1998).
Fig. 1.
The four general pathways of membrane vesicle biogenesis. 1 Exosomes arise from an endocytic pathway that begins with the invagination of receptor-coated plasma membrane to form an endosome (endocytic receptors are depicted in purple). 2 Intraluminal vesicles bud off into the endosome, passively or actively incorporating bioactive molecules as they do so. 3 The endosome matures into a MVB, which is subsequently destined for either degradation within a lysosome, or 4 exocytosis whereby exosomal EVs are released into the extracellular milieu. 5 Microvesicles (shedding vesicles) arise from direct budding and fission of portions of the plasma membrane, encapsulating a cargo of cytoplasmic proteins (depicted in yellow) and nucleic acids from the cytosol as they do so. Variables such as the nature or pathological state of the parent cell will influence the type and contents of EVs. 6 The shrinkage and fragmentation of apoptotic cells gives rise to so-called apoptotic bodies or blebs, 7 while an unknown mechanism believed to involve transcription of endogenous retroviruses leads to the formation of RLPs. (Color figure online)
Exosomes
First identified by Rose Johnstone as a part of the reticulocyte maturation (Johnstone et al. 1987), these EVs were described as being secreted to remove membranes and proteins in a process of reverse endocytosis, and for this reason called exosomes. The biogenesis and trafficking of exosomes is not fully understood. They originate with the invagination of clathrin-coated domains on the plasma membrane, and then enter the cell to be developed by the endosomal network, a membranous compartment that sorts vesicles towards their appropriate sub-cellular destination. The endosomal sorting complex required for transport (ESCRT) machinery (Simons and Raposo 2009; Baietti et al. 2012) is required for transport into early endosomes. Subsequent budding of intraluminal vesicles into the endosomes themselves results in the maturation of the complex into large MVBs. These MVBs are ultimately trafficked to lysosomes for degradation (degradative MVBs) or they fuse with the plasma membrane of the cell (exocytic MVBs), releasing their intraluminal vesicles, at this stage termed exosomes, into the extracellular space.
These latter passages seem to be ESCRT-independent and are instead governed by the distribution of the sphingolipid ceramide and a tetraspannin tertiary structure within raft-based microdomains on the MVB (Trajkovic et al. 2008). This process accounts for the enrichment of ceramide (among other specific lipids and proteins derived from the MVB membrane) observed in exosomes, and also for the abundance of endosome-associated proteins such as Alix and TSG101 (Théry et al. 2002b). However, the relative importance of the ESCRT-dependent or -independent mechanisms is not yet fully elucidated. While the fusion of MVBs with the plasma membrane responsible for the release of exosomes into the extracellular space is reportedly controlled by Rab GTPases (Hsu et al. 2010; Ostrowski et al. 2010), recently an alternative mechanism for the secretion of Wingless-related integration site (Wnt)-bound exosomes was proposed involving the R-SNARE protein YKT6 (Gross et al. 2012).
Microvesicles
The mechanism behind the generation of microvesicles is largely unknown. They represent a heterogeneous population of vesicles that are formed by the outward budding and fission of the cell membrane. Secretion of shedding vesicles may be controlled by cholesterol-rich lipid rafts in the plasma membrane (Del Conde et al. 2005). Moreover, the asymmetric distribution of proteins and phospholipids is tightly regulated by aminophospholipid translocases (Zwaal and Schroit 1997; Bevers et al. 1999; Leventis and Grinstein 2010). Microvesicle formation is induced by translocation of phosphatidylserine to the outer-membrane leaflet (Zwaal and Schroit 1997; Hugel et al. 2005), and the budding process is completed through contraction of cytoskeletal structures by actin–myosin interactions (McConnell et al. 2009; Muralidharan-Chari et al. 2009), regulated in turn by the small GTPase ADP-ribosylation factor 6 (ARF6) (Muralidharan-Chari et al. 2009). Acid sphingomyelinases have also been implicated in microvesicle secretion, notably in glia following adenosine triphosphate (ATP) stimulation: upon ATP activation of the P2X7 receptor, acid sphingomyelinases relocate to the outer leaflet of the plasma membrane, immediately preceding microparticle and interleukin-1β (IL-1β) secretion (Bianco et al. 2005). Inhibition or knockout of acid sphingomyelinase was found to reduce or block, respectively, ATP-induced secretion. Interestingly, a recent study provided evidence for the recruitment of TSG101, an ESCRT subunit, to the plasma membrane and into microvesicles (Nabhan et al. 2012).
Thus, the molecular machineries for exosome and microvesicle biogenesis may share mechanistic elements.
Retrovirus-like particles (RLPs)
The origin of RLPs is still uncertain. They may arise from transcription of human endogenous retrovirus sequences, which represent approximately 8 % of the human genome but are normally silent. Derepression of such sequences can occur following cellular stress (e.g. cytokine stimulation or cancer) (Depil et al. 2002; Reiche et al. 2010; Golan et al. 2008; Wang-Johanning et al. 2003; Taruscio and Mantovani 2004). RLPs arise by directly budding from the plasma membrane with a mechanism involving the interaction of retroviral proteins (i.e. Gag) with components of the plasma membrane (Bieda et al. 2001; Pincetic and Leis 2009) and the cytoskeleton (Gladnikoff et al. 2009). However, their biogenesis is thought to be distinct, even if the size overlaps with exosomes and makes difficult the differential purification.
Apoptotic bodies
Whereas other EVs are secreted during physiological cellular processes, apoptotic bodies arise only during programmed cell death. Like shedding vesicles, a flipflopping process during vesicle blebbing results in high levels of phosphatidylserine on their outer surface. These translocated phosphatidylserines bind to Annexin V, which is subsequently recognized by macrophages for phagocytic clearance (Martinez and Freyssinet 2001).
Thus, elucidation of the mechanisms that give rise to the various types of EV (possibly hindered by an unwieldy and inconsistent designation system) is still far from complete. Only with a full knowledge of the molecular machineries required for the EV biogenesis will researchers be able to thoroughly illuminate the specific origins of each class of EV, and to resolve their respective functions.
The functions of EVs
The content of EVs
EVs contain a broad range of molecules, primarily RNAs, proteins and lipids; according to Vesiclepedia (Kalra et al. 2012), a manually curated database of EV contents, 43,731 different proteins, 20,196 different mRNAs, 2,400 different microRNAs (miRNAs) and 342 different lipids have been described at least once within EVs (database accessed 30th Jan 2013). Some of these are found in most EVs, or are specific markers for a particular EV class, while other vary according to the organism, organ, cell-type and condition of the cell of origin (Théry et al. 2009).
Taking the example of exosomes, trafficked proteins include the numerous components of the endosomal compartment, such as proteins involved in membrane transport, tetraspannins (e.g. CD9, CD63, CD81), MVB proteins (Alix, Tsg101) and Heat Shock Proteins (e.g. Hsp90).
In addition to proteins, evidence is available that several classes of RNAs can be profiled within exosomes. These include mRNAs, miRNAs, viral RNAs and other non-coding RNAs (ncRNAs) (Belting and Wittrup 2008; Janowska-Wieczorek et al. 2005; Nguyen et al. 2003; Skog et al. 2008; Valadi et al. 2007; Zomer et al. 2010), and in some cases exosomal RNAs have been shown to be intact and functional by means of in vitro translation (Valadi et al. 2007). In 2006, Ratajczak et al. demonstrated that EVs derived from embryonic stem cells are enriched in mRNA for several early pluripotent transcription factors capable of reprogramming recipient hematopoietic progenitor cells (Ratajczak et al. 2006). Similarly, EMVs derived from human endothelial progenitor cells were shown to be enriched in a specific subset of cellular mRNAs associated with angiogenic pathways, such as the PI3K/AKT and eNOS signalling pathways, thus potentiating them towards triggering an angiogenic program in target endothelial cells (Deregibus et al. 2007). The functional transfer of miRNAs has been demonstrated by Montecalvo et al., who showed that exosomal miR-148a, abundant in exosomes from bone marrow-derived dendritic cells (DCs), could downregulate an artificially-induced miR-148a target sequence in a miR-148-deficient DC2.4 dendritic cell (DC) line (Montecalvo et al. 2012).
In some instances, the repertoire of proteins and RNAs contained within EVs matches closely that of the cell of origin. However, it has also been found that extracellular signalling is able to modulate the RNA and protein content of EVs. For example, it was shown that stress conditions such as hypoxia alter the protein and RNA composition of exosomes derived from endothelial cells (de Jong and Verhaar 2012). Levels of the mRNAs N-myc downstream regulated 1 (NDRG1) and BCL2/adenovirus E1B 19 kDa interacting protein 3 (BNIP3), stress and apoptosis-related respectively, were significantly upregulated in exosomes from hypoxic cells, whereas cold inducible RNA binding protein (CIRP) mRNA was downregulated. The array of proteins overexpressed in these same exosomes includes lysyl oxidase homolog 2, fibronectin and collagen, suggesting a role in cyto-skeletal and extracellular matrix rearrangement. Furthermore, stress conditions, such as heat stress, oxidative stress, or hypoxia, induce the exosomal secretion of heat-shock proteins (HSPs) in several cell types (Clayton et al. 2005; Eldh et al. 2010; Gastpar et al. 2005; Gupta and Knowlton 2007; Lancaster and Febbraio 2005; Taylor et al. 2007; Zhan et al. 2009). Similarly, it has been shown that the content of exosomes changes under a diversity of conditions: reticulocyte activation induces changes in proteolipidic composition (Carayon et al. 2011); viral infection results in the trafficking of viral miRNAs, such as the secretion of immunosuppressive miRNAs by Epstein–Barr virus infected B cells (Pegtel et al. 2010); and in response to signalling pathway activation, with proteins such as maspin, cyclophilin A, and phosphoglycerate kinase 1 upregulated in exosomes in a p53-dependent manner (Yu et al. 2006). All these evidences point in the direction that there is a cellular machinery able to sort specific proteins and/or RNAs towards exosomes. As such, a recent work by Villarroya-Beltri et al. has showed that the heterogeneous nuclear ribonucleoprotein A2B1 (hnRNPA2B1) specifically binds to a 4-nucleotide motif present in a subset of miRNAs and mediates their loading into exosomes (Villarroya-Beltri and Gutiérrez-Vázquez 2013), reinforcing the idea the exosomal cargo is the result of an active and regulated process. While exosomal miRNA-loading was found to be modulated by changes in hnRNPA2B1 expression, how extrinsic factors might influence this process is still unknown. Moreover, hnRNPA2B1 is also implicated in intra-cellular trafficking and localisation of specific mRNAs in neurons (Munro et al. 1999) and HIV genomic RNA (Levesque et al. 2006), however the role, if any, of hnRNPA2B1 in loading mRNAs into EVs remains to be elucidated.
Mechanism of EV-mediated cell-to-cell communication
According to the type of EV and to the biological context, different mechanisms of interaction between EVs and target cells have been described, including ligand–receptor interactions, internalisation and direct membrane fusion.
DC-derived exosomes containing MHC-peptide complexes are efficiently recruited by T cell and mediate T cell inhibition without being internalised or fusing with the plasma membrane (Nolte-’t Hoen et al. 2009), providing an example of ligand-receptor interactions. Alternatively, other works have shown that EVs, and in particular exosomes, can also be internalised by target cells via endocytosis and macropinocytosis. For example, circulating exosomes are taken up by DCs, phagocytes of the spleen and Kupffer cells in the liver via clathrin-dependent endocytosis (Morelli et al. 2004), while exosomes secreted by oligodendrocytes can be internalised by microglia via macropinocytosis (Fitzner et al. 2011). Also within the brain, Frühbeis et al. showed that glutamate triggers the release of exosomes from oligodendrocytes, the secretion of which is modulated by Ca2+ uptake by N-methyl-d-aspartate (NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors (Fruhbeis et al. 2013b). Exosomes also play an important role in the signalling between oligodendrocytes and neurons, potentially contributing to the myelination and long-term axonal survival of the latter. Similarly, dedifferentiated Schwann cells are found to secrete exosomes, which are taken up selectively by dorsal root ganglia axons, enhancing regeneration in injury models (Lopez-Verrilli et al. 2013). An additional mechanism of EV uptake that has been described is the direct fusion of the vesicle with the plasma membrane (Del Conde et al. 2005), while a recent work by Christianson et al. showed that heparan sulfate proteoglycans (HSPGs) act as receptors for cancer cell-derived exosomes and are required for their internalisation and function in the target cells (Christianson et al. 2013).
EVs as intercellular mediators of physiology and pathology
Recent works have also started to shed light on the function of EVs in physiology and pathology. One of the earliest insights into EV function dates back to the 1980s, when Johnstone and colleagues described EV secretion by sheep reticulocytes, suggesting that it could be a mechanism of protein clearance during reticulocyte maturation (Johnstone et al. 1987). More recently, most of the research efforts on EV function shifted towards immunology and immunotherapy. In a pioneering work Raposo et al. have shown that B cell-derived exosomes are capable of modulating the immune response by spreading MHC-antigen complexes (Raposo et al. 1996). Subsequently, it was shown that the injection of antigen-bearing exosomes derived from DCs induces the activation of antigen-specific CD4+ T cells in vivo, causing an amplification of the primary immune response (Théry et al. 2002a). Additionally, the EV-mediated transfer of MHC-peptide complexes between DCs and from DCs to T cells enhances T cell activation in vitro (Nolte-’t Hoen et al. 2009; Arnold and Mannie 1999; Bedford et al. 1999; Patel et al. 1999). On the other hand, there are also several works reporting that EVs—and in particular tumour derived exosomes—have an immunosuppressive effect in vitro on T cells and NK cells, and promote the induction of T regulatory cells (Clayton et al. 2007; Andreola et al. 2002; Huber et al. 2005; Szajnik et al. 2010; Zeelenberg et al. 2008; Valenti et al. 2006; Liu et al. 2006). It has been shown that bioactive Fas ligand (FasL) and TNF-related apoptosis inducing ligand (TRAIL) are expressed in tumour-derived exosomes and induce apoptosis in activated tumour-specific T cells (Iero et al. 2007), while NK cells lose their cytolytic potential through an exosome-mediated inhibition of perforin release (Liu et al. 2006). Furthermore, tumour-derived exosomes are known to impair the capacity of CD14+ monocytes to differentiate into functional DCs, leading to an abundance of CD14+ cells with low levels of expressed human leukocyte antigen-DR (HLA-DR) that serve as myeloid suppressor cells (Valenti et al. 2006). These data support the idea that tumour-derived exosomes might induce immune tolerance and contribute to tumour growth.
Similarly, others have described that several other cell-types also secrete exosomes carrying immune-suppressive agents. For example, exosomes derived from the placenta carry immunosuppressive FasL and UL-16 binding proteins that modulate the activity of maternal cytotoxic T and NK cell, respectively, inducing tolerance toward the foetus (Hedlund et al. 2009; Taylor et al. 2006). Given the amount of evidence supporting both the immune-stimulatory and immune-suppressive role of EVs, their effect is probably very much dependent on the cell of origin (and therefore the content of the EV), on the state of the target cell, and on the biological context in which the interaction between EVs and target cells takes place.
In addition to their immune-modulatory effect, EVs were also shown to be involved in cytokine activity. For example, exosome-like vesicles mediate the release of full-length TNF receptor 1 (Hawari et al. 2004) and are considered a major mechanism through which murine bone marrow derived macrophages (BMDM) secrete Interleukin-1β (IL-1β) (Qu et al. 2007). A similar mechanism has been observed in microglia following stimulation by astrocyte-derived ATP (Bianco et al. 2005). Additionally, EVs mediate the transfer of the chemokine receptor CCR5 from peripheral blood mononuclear cells to cells that do not express it. The efficient infection of cells by the human immunodeficiency virus-1 (HIV-1) requires the presence of CD4 and a specific chemokine co-receptor, a role served by CCR5 in macrophage-tropic (M-tropic) HIV-1 strains. M-tropic HIV-1 is known to infect multiple cell types, however the expression of CCR5 by target endothelial cells, astrocytes and renal cells is still debated, raising the question as to how such cells become infected. Mack et al. report a potential explanatory mechanism whereby functional CCR5 is transferred via EVs to endothelial cells that do not normally express CCR5 (Mack et al. 2000).
In the last few years it was also reported that EVs can mediate the spread of infections. Wiley and Gummuluru have shown that HIV-1 particles can be endocytosed by DCs and released within exosomes, which in turn can spread the infection to T cells with an efficiency 10-fold higher than cell-free viral particles (Wiley and Gummuluru 2006). Similarly, other works have shown that prion-infected cells release the prion protein (PrPC) and its abnormally folded version scrapie (PrPSc) inside exosomes, which in turn are able to spread the infection to other cells (Fevrier et al. 2004). In addition to the direct spreading of infections there’s also evidence that some pathogens can exploit EVs to modulate their host. For example, Epstein–Barr Virus (EBV)-infected B cells release exosomes that contain viral microRNAs, which in turn are transferred to non-infected cells of the immune system (Pegtel et al. 2010), making this a mechanism by which viruses could potentially modulate the immune response of the infected organism.
As well as being exploited for the spreading of viral infections, EVs were also shown to mediate the intercellular transfer of anti-viral activity. In fact, Li et al. recently showed that IFN-α stimulation of macrophages and liver sinusoidal endothelial cells induces the secretion of exosomes that block Hepatitis B Virus (HBV) replication in infected cells (Li et al. 2013a). This observation suggests the existence of a mechanism that allows uninfected cells to overcome the HBV-mediated blockage of IFN activity in infected cells.
In addition to their role in infections and immunity, EVs can also mediate the acquisition of new functional properties by recipient cells, such as migratory, adhesive or metastatic abilities. For example, in gliomas EVs mediate the transfer of an oncogenic, truncated form of the epidermal growth factor receptor (EGFR) to cells that do not express it, and thus they promote the activation of transforming signalling pathways inducing morphological transformation and promoting growth (Al-Nedawi et al. 2008). Following the discovery that tumour-derived EVs contain onco-genes, other groups have investigated the possibility of using EVs as biomarkers. For example, Skog et al. have found that EVs purified from the serum of glioblastoma patients contain the mRNA for the oncogenic form of EGFR (EGFRvIII) highlighting their potential as diagnostic markers (Skog et al. 2008). In addition to glioblastoma, the diagnostic potential of exosomes is under investigation also for prostate cancer, with various studies having identified altered levels of specific miRNAs in exosomes derived from the serum of prostate cancer patients (Hessvik et al. 2013; Brase et al. 2011; Lodes et al. 2009; Mitchell et al. 2008; Moltzahn et al. 2011). Study into the pathological role of EV-mediated miRNA transfer, and their potential application as disease biomarkers or even therapeutic agents, is a burgeoning field of interest and many potential targets have been identified (see Table 1).
Table 1.
A selection of miRNAs identified in EVs as relevant to pathophysiological conditions or putative therapeutic applications
Disease | EV source | miRNA contenta | Putative effect/target | Reference |
---|---|---|---|---|
Pathological roles of EV miRNAs | ||||
HIV-associated neurodegeneration | Astrocytes treated with pathogenic HIV Tat protein and morphine | miR-29b | Decreased PDGF-B expression and viability in recipient neurons | Hu et al. (2012) |
Prion disease | Prion-infected neuronal cells | let-7b, let-7i, miR-128a, miR-21, miR-222, miR-29b, miR-342-3p, miR-424 (miR-146a) | Increased expression of the cellular prion protein gene | Bellingham et al. (2012a, b) |
Asthma | Bronchoalveolar lavage fluid | let-7 and miR-200 families | Biomarkers of mild nonsymptomatic asthma | Levanen et al. (2013) |
Cardiovascular disease | Injured/dying cardiomyocytes | miR-133a | Biomarker; regulation of cardiac hypertrophy | Kuwabara et al. (2011) |
Urine | miR-4516, miR-3183, (miR-3940-5p), (miR-4649-5p) | Biomarkers of salt sensitivity or inverse salt sensitivity index | Gildea et al. (2013) | |
Plasma of atherosclerosis patients | miR-150 | Reduced c-Myb expression and increased cell migration in recipient microvascular endothelial cells | Zhang et al. (2010) | |
Liver disease | Serum/plasma | miR-122, miR-155 | Biomarkers of alcoholic liver disease and inflammatory liver injury | Bala et al. (2012) |
Kidney disease | Urine | (miR-29c) | Biomarker correlating with renal function and degree of histological fibrosis | Lv et al. (2013) |
Sjögren’s syndrome | Glandular saliva | miR-23a* | Diagnostic/biomarker, salivary gland pathologies | Michael et al. (2010) |
Metabolic diseases | Large adipocytes | miR-16, miR-27a, miR-146b, miR-222 | Stimulation of lipid synthesis, lipid droplet biogenesis and cell growth in recipient small adipocytes | Mueller et al. (2011) |
Blood | miR-130a, miR-195 | Biomarkers of hypertension | Karolina et al. (2012) | |
miR-197, miR-23a, miR-509-5p | Biomarkers of hypercholesterolemia | |||
miR-27a, miR-320a | Biomarkers of type 2 diabetes | |||
Schizophrenia | Post-mortem prefrontal cortices | miR-497 | Putative biomarker; pathogenesis of neoplasms, neurodegenerative diseases and heart disease; promotion of ischemic neuronal death by negatively regulating anti-apoptotic proteins, bcl-2 and bcl-w | Banigan et al. (2013) |
Bipolar disorder | Post-mortem prefrontal cortices | miR-29c | Putative biomarker; regulation of cell-adhesion machinery components | Banigan et al. (2013) |
Colorectal cancer | Serum of colorectal cancer patients | let-7a, miR-1229, miR-1246, miR-150, miR-21, miR-223, and miR-23a | Biomarkers | Ogata-Kawata et al. (2014) |
Plasma of mice with colorectal cancer xenografts | miR-92a | Enhanced proliferation and motility in recipient endothelial cells, down-regulation of target Dickkopf-3 tumor-suppressive gene | Yamada et al. (2013) | |
Melanoma | A375 melanoma cells | let-7a, miR-182, miR-221, miR-222, miR-31, miR-19b-2, miR-20b and miR-92a-2, miR-21, miR-15b, miR-210, miR-30b, miR-30d, miR-532-5p, miR-185 | Melanoma progression and metastasis | Xiao et al. (2012) |
Glioma | Primary human glioblastoma cells | miR-21, let-7a, miR-16 | Tumorigenesis, promotes angiogenesis in target human brain microvascular endothelial cells | Skog et al. (2008) |
Liver cancer | Hepatocellular carcinoma cells | miR-584, miR-517c, miR-378, miR-520f, miR-142-5p, miR-451, miR-518d, miR-215, miR-376a*, miR-133b, miR-367 | Downregulation of transforming growth factor b activated kinase-1 (TAK1) in recipient cells, promotion of hepatocarcinogenesis | Kogure et al. (2011) |
Kidney cancer | Renal cell carcinoma | miR-29a, miR-650, miR-151, miR-19b, miR-29c, miR-200c, miR-92, miR-141 | Induction of an activated angiogenic phenotype in recipient endothelial cells, formation of a pre-metastatic niche | Grange et al. (2011) |
Ovarian cancer | Ovarian tumor cells | miR-21, miR-141, miR-200a, miR-200b, miR-200c, miR-203, miR-205, miR-214 | Biomarkers, tumorigenesis | Taylor and Gercel-Taylor (2008) |
Ovarian cancer cell lines | let-7 family | Levels indicative of parent cell invasiveness | Kobayashi et al. (2014) | |
Prostate cancer | Plasma and serum from prostate cancer patients | miR-375, miR-141 | Metastasis | Bryant et al. (2012) |
miR-107, miR-574-3p | Diagnostic biomarkers | |||
Lung cancer | Plasma from NSCLC patients | (let-7f), (miR-20b), (miR-30e-3) | NSCLC diagnosis and prognosis | Silva et al. (2011) |
Plasma from lung SCC patients | miR-205, miR-19a, miR-19b, miR -30b, miR-20a | SCC biomarkers; oncomiRs | Aushev et al. (2013) | |
Tumorigenicity of Epstein–Barr Virus (EBV) | Nasopharyngeal carcinoma cells | EBV miRNAs | Angiogenesis, cell proliferation, tumor-cell invasion, and immune evasion in recipient human umbilical vein endothelial cells | Meckes et al. (2010) |
Gastric cancer | AZ-P7a metastatic gastric cancer cells | let-7 family | Depletion of tumor-suppressive let-7 miRNA in parent gastric cancer cells, maintaining oncogenesis | Ohshima et al. (2010) |
Gastric cancer tissue-derived mesenchymal stem cells | miR-221 | Promotes proliferation and migration in recipient human gastric cancer cells | Wang et al. (2014a, b) | |
Breast cancer | Activated tumour-associated macrophages | miR-223 | Increases invasiveness of co-cultured breast cancer cells | Yang et al. (2011) |
Breast cancer cell lines | miR-210 | Enhanced angiogenesis and induction of a metastatic niche in recipient endothelial cells | Kosaka et al. (2013) | |
MDA-MB 231 breast cancer cell line | miR-130a | Tumorigenesis through regulation of TGF-β/Smad signalling | Kruger et al. (2014) | |
miR-328 | Targets CD44, reducing cell adhesion, enhancing cell migration, and regulating the formation of capillary structure | |||
MCF-7 breast cancer cell line | miR-301a | Considered a negative prognostic indicator in lymph node negative invasive ductal breast cancer | ||
miR-34a | p53 regulation | |||
miR-106b | Downregulation of BRMS1 and RB, promoting breast cancer invasion and metastasis; mediation of TGF-β-induced epithelial-mesenchymal transfer, an early process in tumor metastasis | |||
Malignant mammalian epithelial cells | miR-451 | Tumour suppression (sequestration thereof); proliferation; cell polarity; dysregulation of oncogenic pathways | Palma et al. (2012) | |
miR-1246 | Induction of p53 dependent apoptosis | |||
Cancer (angiogenesis/metastasis) | Hypoxic K562 leukaemia cell lines | miR-210 | Induction of angiogenesis in recipient endothelial cells | Tadokoro et al. (2013) |
Brain metastatic cancer cell lines | miR-210, (miR-19a), (miR-29c) | Biomarkers; potential prognostic agent for brain metastatic breast cancer and melanoma | Camacho et al. (2013) | |
Metastatic rat adenocarcinoma | miR-494, miR-542-3p | Downregulation of cadhedrin-17 and concomitant up-regulation of matrix metalloproteinase transcription, preparation of a pre-metastatic niche | Rana et al. (2013) | |
K562 leukaemia cells | miR-92a | Reduced expression of integrin α5 in recipient HUVECs, enhancing cell migration and tube formation | Umezu et al. (2013) | |
Therapeutic potential of EV miRNAs | ||||
Cancer (angiogenesis) | MSCs | miR-16 | VEGF down-regulation in recipient tumour cells, anti-angiogenic effect | Lee et al. (2013) |
Highly metastatic A549 adenocarcinoma subpopulation enriched in miR-192 | miR-192 | Repression of pro-angiogenic IL-8, ICAM and CXCL1 in HUVEC co-cultures in vitro, impairs tumour-induced angiogenesis in bone metastasis in vivo | Valencia et al. (2014) | |
Glioma | Marrow stromal cells engineered to overexpress miR-146b | miR-146b | Reduced glioma xenograft growth | Katakowski et al. (2013) |
Breast cancer | Human embryonic kidney cell line engineered to express EGFR-binding peptide, lipofected with let-7a | let-7a | Inhibited tumour development in vivo (xenografted breast cancer cells) | Ohno et al. (2013) |
Prostate cancer | COS-7 fibroblast-like kidney cells transduced with miR-146a | miR-146a | Knockdown of ROCK1 in recipient PC-3 M metastatic prostate cancer cells, attenuating proliferation | Kosaka et al. 2010 |
Epithelial prostate PNT-2 cells | miR-143 | Knockdown of KRAS and ERK5 in recipient PC-3 M metastatic prostate cancer cells, attenuating proliferation | Kosaka et al. (2012) | |
Stroke | MSCs | miR-133b | Enhanced neurite remodelling; Increased functional recovery and neurovascular plasticity in rat stroke models | Xin et al., 2012 |
Xin et al. (2013a), Xin et al. (2013b) | ||||
Multiple sclerosis | Sera of young rats or rats exposed to environmental enrichment | miR-219 | Myelin production, oligodendrocyte differentiation | Pusic and Kraig 2014 |
IFN-γ stimulated DCs | miR-219 | Myelin production, oligodendrocyte differentiation | Pusic et al. (2014) | |
miR-106a, miR-124, miR-181a, miR-451, miR-532-5p, miR-665 | Anti-inflammatory response | |||
Atherosclerosis | KLF2-transduced or shear-stressed HUVECs | miR-143/145 cluster | Regulation of smooth muscle cell function, reduced atherosclerotic lesion formation | Hergenreider et al. 2012 |
Kidney disease | Endothelial progenitor cells | miR-126, miR-296 | Protective effects in models of ischemia–reperfusion injury | Cantaluppi et al. 2012 |
BRMS1 breast cancer metastasis suppressor 1; DC dendritic cell; EBV Epstein–Barr Virus; EGFR epidermal growth factor receptor; HIV human immunodeficiency virus; ERK5 extracellular-signal-regulated kinase 5; IFN interferon; HUVEC human umbilical vein endothelial cell; KLF2 Krüppel-like factor 2; KRAS V-Ki-ras2 Kirsten rat sarcoma viral oncogene homolog; MSC mesenchymal stem cell; NSCLC non-small cell lung cancer; PDGF-B platelet-derived growth factor beta; RB retinoblastoma; ROCK1 Rho-associated, coiled-coil containing protein kinase 1; SCC squamous cell carcinoma; TGF transforming growth factor; VEGF vascular endothelial growth factor
miRNAs enriched into EVs relative to non-pathological conditions; miRNAs in parentheses are depleted
In parallel, numerous works have identified distinct proteins, mRNAs and lipids in exosomes purified from blood or urine of prostate cancer patients, offering additional possibilities for the use of exosomes as disease biomarkers or indicators of treatment efficacy (Soekmadji et al. 2013).
EVs in the brain
The physiological processes of the brain require a highly complex array of intercellular communications between a diversity of cell types over variable distances and time-scales. Mechanisms implicated in neural communications networks include the development of gap junctions, cell adhesion processes, and the secretion of bioactive signalling molecules, neurotransmitters and growth factors. In recent years, mounting evidence has implicated EVs as an additional route of communication within the brain (and, by extension, the broader CNS) (Sharma et al. 2013; Lai and Breakefield 2012; Von Bartheld and Altick 2011). EV secretion has been observed in nearly all cell types that constitute the brain: neurons (Faure et al. 2006; Putz et al. 2008; Schiera et al. 2007), astrocytes (Guescini et al. 2010; Taylor et al. 2007), Schwann cells (Lopez-Verrilli et al. 2013; Lopez-Verrilli and Court 2012), neural stem/progenitor cells (Huttner et al. 2008; Marzesco et al. 2005), microglia (Bianco et al. 2005; Bianco et al. 2009; Potolicchio et al. 2005; Tamboli et al. 2010), oligodendrocytes (Fitzner et al. 2011; Hsu et al. 2010; Trajkovic et al. 2008) and endothelial cells (Simak et al. 2006; Jung et al. 2009) can communicate with each other within the brain and, by extension, the broader CNS.
EVs released from neurons have been confirmed as being involved in synaptic function (Faure et al. 2006), with their release being stimulated by enhanced glutamatergic activity and resulting in increased spontaneous neuronal activity with the presence of glutamate receptor 2 subunits in EVs (Lachenal et al. 2011). Furthermore, EVs have been shown to regulate the synaptic transfer of Wnt morphogens at the neuromuscular junction (Korkut et al. 2009), hinting at a potential role in broader Wnt-mediated developmental processes. Additionally, EVs are implicated in various mechanisms within the specialised immune system of the brain (Cossetti et al. 2012). Microglia, prime components of the intrinsic brain immune response, secrete EVs exhibiting MHC class II molecules, the expression of which is upregulated upon stimulation with interferon (IFN)-γ (Potolicchio et al. 2005). Thus, they may reflect the non-professional antigen-presenting activity of their progenitor cells. Moreover, EVs of microglial origin are found to propagate inflammatory signals in vitro and in vivo, with cerebrospinal fluid (CSF) levels of myeloid EVs exhibiting a positive correlation with neurodegenerative disease activity. Mice in which EV secretion had been inhibited showed protection against experimental encephalomyelitis, an animal model of multiple sclerosis (MS) (Verderio et al. 2012), and significantly elevated levels of neurotoxic myeloid EVs have been detected in the CSF of Alzheimer’s disease (AD) patients (Joshi et al. 2014). These observations highlight the role of microglial EVs as not only markers of neuroinflammation, but also putative therapeutic targets for the treatment of neurodegenerative disease. Like microglial EVs, endothelial cell-derived EV levels are particularly responsive to the immune state of the CNS, with increased secretion under inflammatory conditions making them putative biomarkers of cerebrovascular disorders and neuroinflammatory diseases such as MS (Minagar et al. 2001). Levels of circulating endothelial EVs are being correlated with the severity and prognostic outlook of disease (Simak et al. 2006; Jung et al. 2009), and the EVs themselves are being attributed possible roles in the propagation of inflammation (Chironi et al. 2009; Morel et al. 2011) by stimulating the trans-endothelial migration of monocytes (Jy et al. 2004). Such migration is believed to be facilitated by the binding to and activation of monocytes via a specific phenotypic subset of CD54+ EVs.
Nevertheless, EVs are also believed to serve a protective role with respect to brain injury and regeneration. Trophic support for neurons by oligodendrocytes has been ascribed to exosome-mediated transfer of genuine myelin proteins and stress-protective proteins (Kramer-Albers et al. 2007). Conversely, the neuron-modulated release of auto inhibitory oligodendrocyte-derived exosomes have also been implicated in the inhibition of myelin membrane sheath formation (Bakhti et al. 2011), The cargoes carried by oligodendroglial exosomes, including metabolites, protective proteins, glycolytic enzymes, mRNA and miRNA may serve to maintain axonal integrity (Fruhbeis et al. 2013a). Neuron-derived exosomes are also attributed a role in the sequestration of unwanted Nedd4-family metal cation-transporting proteins during times of stress (Putz et al. 2008), while endothelial cell and astrocyte-derived microvesicles have been found to be enriched in nucleoside triphosphate diphosphohydrolases, imbuing them with the capacity to suppress toxic levels of ATP after an ischemia-related breach of the blood brain barrier (Ceruti et al. 2011). EVs are also found to be a means of degradation of toxic β-amyloid (Aβ) protein, the accumulation of which is implicated as a causative factor in AD, when taken up by microglia; however, pathologic accumulation of Aβ neurons restarts when that clearance pathway is overwhelmed (Yuyama et al. 2012) and EVs are considered to be putative vehicles by which toxic protein aggregates are spread in several neurodegenerative diseases (see below).
EVs in biogerontology
Senescence
With EVs having been attributed a role in intercellular communication, it stands to reason that they too play a significant part in the propagation of senescence/aging-related processes. Indeed, while the study of the role of EVs in aging (at the cellular or organismal level) is still in its infancy, there is evidence that senescent cells do undergo specific changes in EV trafficking, particularly with regards to exosome trafficking.
Cellular senescence, induced by triggers such as shortening of the telomeres, commonly operates via tumour-suppression pathways, notably the p53 pathway. Upon activation, p53 up-regulates secreted factors such as insulin-like growth factor-binding protein 3 (IGFBP-3, a growth factor regulator), maspin (Yu et al. 2006) and plasminogen activator inhibitor 1 (PAI-1, inhibitors of protease activity in the extracellular matrix), and TSP (an antiangiogenic), all modulators of the microenvironment. Moreover, p53 is known to regulate the transcription of several genes involved in the biogenesis and secretion of exosomes, enhancing the extracellular release of exosomes upon senescence-correlated activation (Yu et al. 2006; Yu et al. 2009). p53 is reported to up-regulate the expression of caveolin-1 and charged MVP protein 4C (CHMP4C), two genes involved in the regulation of the endosomal compartment (Yu et al. 2009; Feng 2010). Caveolin-1, the primary component of caveolae plasma membranes, facilitates endocytosis and internalisation of surface receptors such as EGFR (Yu et al. 2009), while CHMP4C plays a role in MVB formation as a part of the ESCRT-III complex (Saksena et al. 2007). Similarly, p53 stimulates the expression of genes ascribed roles in vesicle secretion, notably tumour suppressor-activated pathway 6 (TSAP6) which has been shown to be integral to competent exosomes release (Lespagnol et al. 2008). Thus, senescence-associated chronic activation of the p53 tumour suppression pathway and the associated up-regulation of an array of auto-, para- and endocrine-acting secreted factors, including exosomes, is implicated in the propagation of the aging phenotype from the cellular to organismal levels. Unfortunately, the contents of senescence-evolved exosomes have yet to be characterised to any significant extent, and therefore the specifics of their function in influencing recipient cells during aging in vivo remains ambiguous. Nevertheless, exosomes have been associated with a number of age-related pathologies, both as diagnostic biomarkers and putative propagators of disease.
Chief amongst these aging-related disorders is cancer. An accumulation of mutations (and diminished genetic repair efficacy) and senescence-induced pro-oncogenic tissue changes during aging results in an exponential increase in the occurrence of cancer in older organisms (Krtolica and Campisi 2002). EVs and their contents are established as biomarkers of a number of cancer types (Vlassov et al. 2012; Principe et al. 2013; Gabriel et al. 2013; Lau et al. 2013; Wittmann and Jäck 2010), as well as being putative agents of tumour cell proliferation and metastasis (Azmi et al. 2013; Simona et al. 2013; Shin-ichiro et al. 2013). Tumour-derived EVs are known to traffic a variety of proteins, mRNAs, miRNAs and metabolites that can promote an oncogenic niche, obstructing immune responses, promoting angiogenesis, and yielding an environment more conducive to tumour cell mobilisation. This age-related increase in cancer risk is manifested largely as an increased incidence of epithelial carcinomas (DePinho 2000), with lung, colon, breast and prostate cancers being responsible for the highest cancer mortalities in the elderly (Cancer Research UK: http://www.cancerresearchuk.org/cancer-info/cancerstats/mortality/age/). However, given the brain-related focus of this review we will focus upon glioma, the most common adult-onset brain tumour (Stoll et al. 2013).
Brain cancer
Glioma increases in incidence with age and exosomes have been implicated in its malignancy. Microvesicles secreted by glioma cells are characteristically enriched in tumour-characteristic miRNAs, and proteins and mRNAs capable of inducing pro-angiogenic phentotypic modulation in target brain endothelial cells and stimulating proliferation in an autocrine manner (Skog et al. 2008). Glioma-derived exosomes were found to contain angiogenin, fibroblast growth factor (FGF)-α, IL-6, IL-8, issue inhibitors of metalloproteinases (TIMP)-1, TIMP-2 and vascular endothelial growth factor (VEGF), angiogenic proteins that are envisioned to exert their biological function on recipient endothelial cells, In addition to angiogenesis, ontology analyses reveal high levels of expression within these glioma EVs of mRNAs involved in cell migration, cell proliferation, immune response and histone modification, all potential avenues through which tumours might modulate their stroma and facilitate growth. As described earlier, these trafficked mRNAs also include the oncogenic variant of EGFR, EGFRvIII, a characteristic bio-marker of some clinically distinct glioblastoma subtypes (Pelloski et al. 2007). Glioma-derived EVs are found to promote oncogenic transformation of neighbouring cells via trafficked EGFRvIII which in turn enhanced angiogenesis through induced VEGF expression and a resultant autocrine stimulation of VEGF receptor 2 (Al-Nedawi et al. 2008), These EVs were are also found shuttle the protein cross-linking enzyme, tissue transglutaminase (tTG), which imbued non-transformed fibroblasts and epithelial cells with cancer-like properties including anchorage-dependent growth and enhanced survival capability (Antonyak et al. 2011). Exosome secretion and tumour aggressiveness are once again seen to increase under hypoxic conditions (Svensson et al. 2011), with an enrichment in exosomal MMP, IL-8, platelet-derived growth factors, caveolin-1 and lysyl oxidase (the increased expression of which have been associated with cancer progression and poor prognosis) relative to normoxic secretions (Kucharzewska et al. 2013). Thus, hypoxic conditions appear to drive cancer cells to modulate their microenvironments via microvesicle secretions, yielding a state more conducive to tumour growth and metastasis (Park et al. 2010).
Neurodegenerative disease
Exosomes are implicated in many facets of neuron-to-neuron, neuron-to-glia, glia-to-glia, and glia-to-neuron communication within the CNS (Fruhbeis et al. 2013b), but from a pathophysiological standpoint they are best characterised as conveyors of inflammatory signals and vehicles by which toxic protein aggregates (or their precursors) are propagated (Schneider and Simons 2013; Vella et al. 2008; Kalani et al. 2013). Cellular and molecular changes that occur during the aging process, such as the accumulation of oxidative damage and diminished adaptive immunity, make the elderly more susceptible to neurodegenerative disease, while impaired neurogenesis limits self-repair (Hung et al. 2010).
AD is the most common form of dementia (Querfurth and LaFerla 2010), and is associated with the accumulation of Aβ peptides into potentially neurotoxic extracellular plaques. The earliest signs that exosomes might be involved in this process came from observations in the 1970s that MVBs were more abundant and larger in cortical dendrites obtained from AD patients (Paula-Barbosa et al. 1978). The Aβ peptide undergoes extensive processing and sub-cellular trafficking, with the amyloidogenic Aβ42 fragment ultimately accumulating in MVBs (Takahashi et al. 2002) and subsequently being secreted into extracellular space via exosomes (Rajendran et al. 2006). Aβ, its parent protein amyloid precursor protein (APP), and β- and γ-secretases (proteases responsible for cleavage of APP into the Aβ peptide) have all been found enriched in exosomes obtained from AD patients, hinting at the possibility that processing of APP into pathogenic forms might be occurring in the exosomal pathway (Vella et al. 2008). Furthermore, exosomal proteins such as Alix and flotillin-1 have been found in association with plaques in the brains of AD patients, implying a potential role of exosomes in the formation of these deposits (Rajendran et al. 2006). Several reports describe the potential role of EV-associated lipids in shifting the equilibrium between monomeric Aβ units, soluble Aβ oligomers, and insoluble Aβ aggregates. Microglia-derived EVs, levels of which are elevated in AD patients, were found to promote the formation of the neurotoxic, soluble oligomeric form of Aβ from insoluble aggregates (Joshi et al. 2014), further implicating EVs in neurodegeneration. On the other hand, neuron- and astrocyte-derived exosomes have been demonstrated to promote aggregation of monomeric Aβ into insoluble plaques, suggesting that the parent cell-dependent lipid composition may influence an EV’s effect on Aβ aggregation (Dinkins et al. 2014; Yuyama et al. 2008). Indeed, exosomes may present a clearance mechanism by which potentially pathogenic deposits are shuttled to microglia for degradation (Yuyama et al. 2012) or degraded by EV-shuttled proteases such as the insulin degrading enzyme (Tamboli et al. 2010). It is also noteworthy that microvesicles isolated from the cerebrospinal fluid of AD patients exhibit some 60 miRNAs that are differentially expressed relative to healthy controls, however the significance of these specific markers has yet to be established (Cogswell et al. 2008). Hyperphosphorylation of the tau micro-tubule-associated protein results in the disruption of tau’s normal axonal transport function as well as the formation of neurofibrillary tangles and toxic species of soluble tau, and these effects have been associated with a number of neurodegenerative diseases, including AD. Indeed, secretion and interneuronal transfer of toxic tau species is believed to play a role in the spread of AD lesions, and selectively phosphorylated tau has been shown to be actively secreted via exosomes into the CSF during early stages of the disease, not just as refuse from dying neurons (Saman et al. 2012).
Parkinson’s disease (PD), the second most common neurodegenerative disease after AD, is also characterised by the accumulation of protein aggregates (Lees et al. 2009; Russo et al. 2012). The pathological progression of the disease, selective degeneration of dopaminergic neurons in the substantia nigra pars compact, is accompanied by the formation of Lewy bodies, deposits primarily consisting of fibrillar α-synuclein (α-Syn), in surviving neurons (Danzer et al. 2012). While the exact mechanism of PD pathogenesis is yet to be elucidated, toxic α-Syn aggregates are implicated and intercellular transport of α-Syn from overexpressing neurons to recipient neuronal cells has been observed (Russo et al. 2012; Danzer et al. 2012). Excess α-Syn secreted by neurons can be phagocytised by astrocytes and microglia in a putative waste clearance mechanism (Lee et al. 2010; Lee et al. 2008), however excessive accumulation in these recipient cells can lead to the formation of inclusions and trigger an inflammatory response (Vekrellis et al. 2011; Lee et al. 2010; Halliday and Stevens 2011). α-Syn can be transferred between neurons, result in aggregation within the recipient neurons, inducing cell death (Desplats et al. 2009; Hansen et al. 2011; Emmanouilidou et al. 2010; Schneider and Simons 2013). Exosomes are known to play a role in this intracellular transfer and propagation of α-Syn (Alvarez-Erviti et al. 2011a; Emmanouilidou et al. 2010; Schneider and Simons 2013) in an active, energy-dependent manner (Bellingham et al. 2012b; Aguzzi and Rajendran 2009). Moreover, it is reported that α-Syn oligomers that are associated with exosomes are more likely to be taken up by recipient cells and are more toxic than free α-Syn (Danzer et al. 2012). PD has been linked to mutations in a number of genes involved in the endosomal-lysosomal pathway, thus it might be speculated that resultant alterations to vesicle secretion and trafficking mechanisms might play a role in disease progression (Russo et al. 2012; Schneider and Simons 2013). One such example is leucine-rich receptor kinase 2 (LRRK2), which is involved in exosome secretion (Alegre-Abarrategui et al. 2009; Dihanich and Manzoni 2011; Shin et al. 2008; Piccoli et al. 2011); mutations to this protein yield abnormally large MVBs which may potentially release large numbers of exosomes bearing α-Syn (Alegre-Abarrategui et al. 2009; Russo et al. 2012).
Like AD and PD, prion disease is associated with misfolded proteins and is more prevalent in the elderly. Indeed, the neuropathology implicated in the various types of prion disease may coexist with the protein aggregates described above (Kovacs and Budka 2002). Prion disease is a fatal, transmissible neurodegenerative disease that involves the conversion of the prion protein PrPC into an abnormal, misfolded and protease-resistant, pathogenic isoform, PrPSc (Brown and Mastrianni 2010). The infectious form of the disease begins in the periphery before spreading to the brain via a yet unknown mechanism (Fevrier et al. 2004). Once acquired, the toxic form of the protein catalyses conversion of the non-toxic form to the pathogenic state, however the means of dissemination of the disease was, until recently, a mystery since no plausible vector by which PrPSc could spread to uninfected tissue could be identified (Vella et al. 2007). Experiments demonstrating that the culture medium of infected cells was itself infectious hinted at an extracellular route of transmission, with subsequent characterisations revealing exosomes to be the likely carriers of the toxic prion (Vella et al. 2007; Alais et al. 2008). PrPSc-laden exosomes from infected cells were found to be of greater density than those from healthy cells, containing only the PrPC form, due to the formation of toxic protein aggregates (Vella et al. 2007). Moreover, exosomes from infected neuronal cells have been described as being more spherical in shape, but diverse in size and internal structure (Coleman et al. 2012), while infected platelets are described as releasing PrPSc via both microvesicles and exosomes (Robertson et al. 2006). In the case of the neuronal cell line, prion packaging into exosomes is believed to involve N-terminal modifications to a distinct subtype of PrP glycoforms (Vella et al. 2007).
Therapeutic applications of EVs
Although the significant and broad role played by EVs has only recently come to receive due attention, and is still far from being thoroughly elucidated, the therapeutic potential of these extracellular delivery vectors is already under intense investigation. Numerous studies have demonstrated the in vivo and in vitro loading of EVs with a diversity of drugs, enzymes, genes and RNAi agents and, furthermore, seen their subsequent application as putative therapeutic vectors in a variety of disease models. Therapeutic applications of EVs are perhaps most promising within the CNS, where conventional drugs have traditionally exhibited low efficacy due to a number of biological barriers to their delivery. The tissue/cell specificity and low immunogenicity of biogenic EVs, coupled with an appropriate cargo of bioactive molecules (be they naturally derived or artificially loaded), makes for a potent therapeutic vector in the treatment of neurodegenerative disease and brain cancers for which age is a prominent risk factor.
Therapeutic potential of EVs in brian repair
Delivery of therapeutic agents into the brain is a challenging task due to the major obstacle of the blood–brain barrier (BBB). Numerous studies have shown the advantages of biological EVs for brain repair (Lakhal and Wood 2011; Zhuang et al. 2011; Alvarez-Erviti et al. 2011b), since they possess the ability to cross BBB, as well as ability to deliver therapeutic cargoes, inherent targeting ability to certain cell types, and immune tolerance. EVs often manifest selective cell homing that, like many key EV features, is often specifically derived from the parent cell. Furthermore, they possess effective protective ability for bioactive cargoes including mRNA, siRNA, miRNA, proteins and drugs, thus making them potential natural vehicles in drug delivery system. Furthermore, advances in genetic engineering allow for the functionalization of otherwise naturally occurring EVs to enhance e.g. their targeting capability or to bolster their therapeutic cargoes. These properties lend themselves to the development of EV-based cell free therapies for brain diseases.
Naturally occurring EVs have innate therapeutic potential due to their diversified bioactive cargoes, thus making them novel candidates as cell-free therapy. Yu et al. describe the isolation of a sub-class of DC-derived exosomes expressing TGF-β1 in their membranes which purportedly exert a potent immunosuppressive effect capable of inhibiting the development and progression of experimental autoimmune encephalomyelitis (EAE) in recipient mice when delivered systemically (Yu et al. 2013). The importance of the host cell type can be seen in the experiments described by Hajrasouliha et al. wherein exosomes obtained from retinal astroglial cells (RACs) were able to suppress retinal vessel leakage and inhibit choroidal neovascularisation, whereas exosomes from retinal pigmental epithelium were not (Hajrasouliha et al. 2013). The anti-angiogenic properties of these RAC-derived EVs were attributed to the exclusive presence of endogenous angiogenesis inhibitors in those exosomes.
There is considerable interest in EV-based RNA-interference (RNAi)-based therapy, with EVs representing an ideal platform for RNA delivery, opening a new route for gene modulation. Recent reports have demonstrated that systemic administration of exosomes derived from mesenchymal stem cells (MSCs) promoted neurovascular remodelling and functional recovery after stroke in rats (Xin et al. 2013a). The authors’ initial hypothesis, that the MSC-derived exosomes were exerting this neurological recovery via transfer of miR-133b, known to enhance neurite remodelling and be at high levels in MSC-derived exosomes (Xin et al. 2012), was supported by follow-up experiments (Xin et al. 2013b). Exosomes were found to transfer miR-133b to neurons and astrocytes, and cause a knockdown in the expression of connective tissue growth factor and ras homolog family member A at the ischemic boundary zone in rat stroke models, enhancing functional recovery. The application of EVs as drug delivery vehicles is another focus of developing EVs as therapeutics. For instance, Sun et al. were the first to devote their efforts to load curcumin, a polyphenol anti-inflammatory compound, into EVs derived from EL-4 lymphoma lines, with this exosomal curcumin affording protection against LPS-induced inflammation in mice through delivery to activated myeloid cells (Sun et al. 2010). Remarkably, the same group reported intranasal administration of exosomal curcumin or exosomal JSI-124, a signal transducer and activator of transcription 3 (Stat3) inhibitor, could across BBB and resulted in suppression of a range of inflammation-driven disease models, including LPS-induced inflammation, myelin oligodendrocyte glycoprotein-induced EAE and GL26 glioma (Zhuang et al. 2011).
Despite the promising properties of naturally occurring EVs, there is considerable interest in improving their therapeutic utility through genetic engineering, with the goals of improving specificity or enrichment in the bioactive cargo(es) of interest. Genetic engineering of EV producer cells or direct modification of the EVs themselves, with insertion of therapeutic agents into the lipid layer or loading into their aqueous core, are proposed as means to modulate the specificity and activity of EVs as targeted delivery vehicles (Lai et al. 2013).
The inspiration for the targeted delivery of EVs was perhaps first envisioned in EV-mediated immunotherapy (Trumpfheller et al. 2012), particularly for cancer vaccines, EVs have been investigated to pulse DCs with antigens to activate an immune response against tumor cells (Tan et al. 2010). Work by Viaud et al. showed that DC-derived highly immunogenic, clinical grade EVs expressing CD40, CD80, CD86, and ICAM-1 on their membranes could prime CD8+ T cells in a peptide-dependent manner (Viaud et al. 2011). By utilizing the ligand-receptor interactions, studies have demonstrated EVs that express ICAM-1 can bind DCs and T cells, while EVs from B cells that carried selected galectins can target T cells (Théry et al. 2009). These encouraging findings laid the foundation for further steps into targeting EVs to brain tumours (Lai and Breakefield 2012) and CNS inflammatory disease. The combinatorial EV-based therapy (El-Andaloussi et al. 2013) that couples DC-derived EVs presenting tumour antigens to T cells with tumour-targeted EVs loaded with RNAi effectors is also expected to be a potent therapeutic approach to CNS diseases.
The first proof-of-concept for applying modified EVs in targeted drug delivery for the brain was from the work done by Alvarez et al., in which the host DCs were engineered to express Lamp2b fused to the neuron-specific peptide rabies virus glycoprotein (RVG), imbuing the daughter EVs with BBB-traversal capabilities and facilitating their subsequent uptake into neurons, microglia and oligodendrocytes. These engineered EVs were able to deliver siRNA into the mouse brain where they achieved strong knockdown of beta-site APP cleaving enzyme 1 (BACE1) mRNA and protein, a therapeutic target for AD (Alvarez-Erviti et al. 2011b). This proposed method, along with identification of targeting peptides selectively binding to the cell type or tissues of interest in the brain, is a significant step towards realising the therapeutic potential of EVs, particularly as RNAi-delivery platforms (El-Andaloussi et al. 2012). In addition, strategies utilised in the modification of artificial nanoparticles, such as utilizing monoclonal antibodies complementary to receptors that are naturally expressed on the BBB (Roberts et al. 1993) or inflamed tissues, could be adopted for EV modification.
Despite the modification of EVs for targeted drug delivery, the target loading of cargoes into EVs is also addressed as an important issue requiring further development. For the loading of nucleic acids, several strategies including transfection-based approaches and electroporation have been utilized. The basic idea of transfection method is to construct suitable expression vectors that can be transfected into donor cells and finally induce overexpression of desired short RNAs enriched into EVs (Kooijmans et al. 2013). Several studies have evidenced successful loading of siRNAs and miRNAs into EVs with constructed vectors as well as utilizing transfection reagents (Olson et al. 2012; Zhang et al. 2010b; Kosaka et al. 2010; Ohno et al. 2013). For instance, in a rat model of primary brain tumour, exosomes derived from MSCs engineered to over-express anti-tumour miR-146b significantly reduced glioma xenograft growth upon intra-tumour injection (Katakowski et al. 2013a). Interestingly, synthetic spherical nucleic (SNAs) acids endocytosed into PC-3 prostate cancer cells were naturally sorted into exosomes to a small degree (<1 %), while transmission electron microscopy results indicated SNAs were internalized into exosomes as well as bound to the membrane surface (Alhasan et al. 2014). Nevertheless, questions remaining in the transfection-based approach revolve around not only the level of desired small RNAs enriched in EVs independent of sequences (Batagov et al. 2011), but also the changing encapsulation process and behaviour of EVs (Kooijmans et al. 2013). As for the electroporation method, loading efficiency may vary among sequences of small RNAs (Kooijmans et al. 2013). Wahlgren et al. showed up to 85.2 % of EVs loaded with exogenous siRNA successfully induced gene knockdown in monocytes or lymphocytes (Wahlgren et al. 2012). Alvarez-Erviti et al. demonstrated that RVG-exosomes loaded with approximately 25 % of the electroporated siRNA induced up to 60 % mRNA and protein knockdown, predominantly in the midbrain, cortex and striatum (Alvarez-Erviti et al. 2011b). However, there is a debate about the efficiency of electroporation. Kooij-mans et al. argued that electroporation is far less efficient than previously described since electroporation of EVs with siRNA is accompanied by extensive siRNA aggregate formation, which may cause overestimation of the amount of siRNA actually loaded into EVs (Kooijmans et al. 2013). Therefore, there is an urgent need to develop efficient approaches to load exogenous cargoes into EVs. Such efforts could be devoted to synthesizing EV-targeted vectors, as well as screening RNA targeting to EVs, since multiple motifs were found specifically enriched in secreted RNAs (Batagov et al. 2011) and a “zip code-like” sequence may direct mRNAs targeting into EVs (Bolukbasi et al. 2012).
Nanotherapeutic synthetic analogues of EVs
The exploitation of natural EVs and their biogenic cargoes as therapeutic agents is a very promising avenue of cellular medicine research. As potent vehicles by which to deliver potentially therapeutic miRNAs and proteins to dysfunctional cells, micro-vesicles may prove invaluable in treating a wide variety of diseases and disorders, including those that are age-associated. Advances in synthetic biology and genetic engineering will further progress our ability to evolve and develop these therapeutic delivery platforms with tailored contents and targeting, but current applications of natural exosomes are limited (Koppers-Lalic et al. 2013; Kosaka et al. 2013; Lai et al. 2013; Munoz et al. 2013; Tan et al. 2013; van Dommelen et al. 2012; Kalani et al. 2013). A number of factors need to be carefully considered in the application of naturally occurring EVs. The composition and contents of the vesicles can be complex and difficult to characterise which may confound predictions of in vivo activity; the abundant and diverse bioactive contents of vesicles may exert a plethora of effects, intended and unintended. Furthermore, obtaining pure microvesicle preparations can be tedious, and scalable production of vesicle from mammalian cells is problematic due to low yields (van Dommelen et al. 2012; Lakhal and Wood 2011). Thus, while advances continue in modulating the expression, composition and contents of natural microvesicles towards more efficacious therapeutic activity, parallel efforts are being made in the development of biomimetic or synthetic drug-delivery platforms (Kooijmans et al. 2012).
There have been a number of novel approaches to generating artificial vesicles that are nonetheless cell-derived, including stem cell nanoghosts and nanovesicles (Jo et al. 2014; Toledano Furman et al. 2013; Jang et al. 2013). However, the recent emergence of nanomedicine, the utilisation of cell- and molecule-specific interactions for medicinal applications, has led to the adoption of a plethora of diverse technologies and synthetic constructs as putative platforms for the cellular delivery of therapeutic and diagnostic agents (Tennyson and Clemens 2012; Duncan and Gaspar 2011; Devadasu et al. 2013; Ganta et al. 2008; Collet et al. 2013; Gao et al. 2013). The makeup of these nanovehicles spans a broad range of physical and chemical compositions (see Fig. 2), including liposomal and polymersomal EMV analogues (Akbarzadeh et al. 2013; Christian et al. 2009; Lee and Feijen 2012; Theresa and Pieter 2013), micelles (Deng et al. 2012; Xu et al. 2013), polymer, protein and lipid complexes (Zia ur et al. 2013; Wasungu and Hoekstra 2006; Ge et al. 2012; Zhang et al. 2012), dendrimers (Deng et al. 2012; Zhu and Shi 2013), and nucleic acid-based nanoparticles and nanostructures (Shu et al. 2014; Roh et al. 2011). Beyond these biomimetic and bioinspired delivery platforms, there has been also considerable interest in the use of surface-functionalised inorganic nanoparticles, nanocrystals, nanotubes and quantum dots for nanomedicinal applications. Such agents are thoroughly reviewed elsewhere (Sekhon and Kamboj 2010a; Malmsten 2013; Rajendra and Hae-Won 2013; Son et al. 2007a, b; Sekhon and Kamboj 2010b) but are beyond the scope of this review. Below we describe some of the synthetic nanoscale drug-delivery systems most reminiscent of biological EMVs, and provide examples of their application in the treatment of age-related disease.
Fig. 2.
Types of synthetic EV analogue nanovehicles. Liposomes (a) are membrane bilayers enclosing an aqueous/hydrophilic interior. Polymersomes (b) are comprised of amphiphilic block copolymers that self-assemble into a sphere with a hydrophobic layer sandwiched between a hydrophilic core and surface. Micelles (c) also consist of amphiphilic block copolymers, but assembled into a sphere with a hydrophobic core and hydrophilic exterior. Polyplexes (d), like their lipid or protein-based analogues, complex polyanionic nucleic acids via electrostatic interactions with cationic polymers. Dendrimers (e) are unimolecular, branched spherical assemblies with dense, hydrophobic surfaces but relatively empty pockets nearer the core in which to encapsulate drugs. (Color figure online)
The nanovehicles most resembling natural micro-vesicles are liposomes and polymersomes. Both are synthetic vesicles of adjustable size (typically tens to hundreds of nanometres in diameter), usually enclosing and protecting an aqueous compartment, however the membrane of the former consists of a lipid bilayer (typically comprised of phospholipids) while the latter is self-assembled from amphiphilic block copolymers (Chandrawati and Caruso 2012; LoPresti et al. 2009; Allen and Cullis 2013). Hydrophilic cargoes are enclosed within the aqueous compartment, whereas hydrophobic species can be sequestered within the membrane; the vesicles, like all drug delivery vehicles, serve as vectors by which to enhance drug pharmacokinetics, uptake, stability or solubility, or as a means to mask the (off-target) toxicity of the cargo. Both classes of vesicle are tailorable in composition, however liposomes generally benefit from a high biocompatibility and a soft and fluid bilayer, which can facilitate direct interaction with cell membranes, whereas polymersomes typically possess a greater mechanical and chemical stability making them potentially more robust in regards to functionalization. While some liposomes can enter cells via direct membrane fusion (dependent on liposome membrane composition) (Lee et al. 2005), many liposomes and most polymersomes are proposed to enter cells via an endocytic pathway; passively targeted vesicles are believed to be taken up through pinocytosis (or phagocytosis) whereas those actively targeted towards a specific cellular surface marker benefit from receptor-mediated endocytosis (Allen and Cullis 2013; Templeton 2002; Christian et al. 2009).
Early liposomes applications involved the use of neutrally-charged vesicles for the delivery of proteins and drugs (Gregoriadis and Ryman 1971), and later genes (Tai-Kin et al. 1980). More recently, cationic PEGylated (that is, coated in bio-inert poly(ethyleneglycol)) liposomes have been employed as the du jour standard for greater transfection ability and biocompatibility (Collet et al. 2013; Boado 2007), and liposomal cargoes have expanded to include RNAi agents (Spagnou et al. 2004; Kanasty et al. 2013; Buyens et al. 2012). Complexes of cationic lipids and polyanionic nucleic acids are sometimes referred to as lipoplexes (Wasungu and Hoekstra 2006; Zhang et al. 2012). Over the years, there has been considerable research into further optimising the pharmacokinetics of liposomes and improving the encapsulation extent, release rate, and intracellular delivery of liposome-delivered therapeutics (Allen and Cullis 2013). Through appropriate modification and functionalization of the lipidic membrane, liposomes can be engineered to release their contents under appropriate trigger conditions, such as a specific pH range, elevated temperatures, irradiation, sonication, or enzymatic degradation (Allen and Cullis 2013). Furthermore, there have been extensive efforts to develop actively targeted liposomes through the attachment of cell-specific ligands, or their incorporation into the lipid formulation, with the intent of enhancing drug delivery to the tissue of interest. This is typically approached by using monoclonal antibodies (mAbs) to direct so-called immunoliposomes against surface receptors on the cells of interest (Noble et al. 2014), however enzymes (Blume et al. 1993), small molecules (Lee and Low 1994), and nucleic acid aptamers (Cao et al. 2009) have also been employed. Nevertheless, the efficacy of targeted liposomes has to date generally not been considered sufficiently improved over passively-targeted liposomes to warrant the extra preparative work and cost (Allen and Cullis 2013).
Liposome-drug formulations have been approved for clinical applications, especially the delivery of anti-cancer chemotherapeutics via a variety of administrative routes, or are presently in clinical trials (Allen and Cullis 2013). Indeed, PEGylated, doxorubicin-loaded liposomes—Doxil—became the first nano-medicine to be approved by the FDA in 1995 (Gabizon et al. 1994; Barenholz 2012). Of specific relevance to age-related diseases, liposome-delivered doxorubicin has also been approved for use in treating early and metastatic breast cancers (Lao et al. 2013), with HER2-targeted (Hendriks et al. 2013) and hyperthermia-triggered (Staruch et al. 2011) liposomes undergoing clinical testing as delivery vectors. Anti-tumour applications are not restricted to the delivery of doxorubicin: for instance, liposomes have been used to deliver paclitaxel to breast cancers (Fasol et al. 2012), cisplatin and anti-MUC vaccines in non-small cell lung cancers (Fantini et al. 2011; Bradbury and Shepherd 2008), irinotecan (CPT-11) to colon and breast cancers (Drummond et al. 2006), and combinatorial treatments (irinotecan and floxuridine) in colorectal cancers (Batist et al. 2008). Overexpression of miR-7 in tumour models through the liposomal delivery of a miR-7 plasmid was found to lead to the suppression of EGFR tyrosine kinase inhibitor-resistance in lung cancer cells (Rai et al. 2011), while immunoliposomes targeted towards human insulin receptor and mouse transferrin receptor (TfR) delivered an anti-EGFR shRNA plasmid, knocking down EGFR expression and increasing survival in murine glioma models (Zhang et al. 2004). Liposome-based treatments for cardiovascular disease are also undergoing clinical trials, with the vesicles being employed as vehicles to deliver RNAi-based therapeutics, such as anti-proprotein convertase subtilisin/kexin type 9 (PCSK9) siRNA for tackling hypercholesterolemia (Jayaraman et al. 2012).
To date, liposomes have received relatively little attention as therapeutic delivery agents for the treatment of neurodegenerative disease. Some success has been achieved with directing liposomes across the BBB using receptor-mediated transcytosis and appropriate mAbs (e.g. the insulin and TfR antibodies mentioned above) (Boado 2007). Therapeutic outcomes were achieved in rats with experimental PD by delivering TfR-targeted liposomes across the BBB and into neurons, delivering a glial-derived neurotrophic factor (GDNF) plasmid. Expression of the GDNF was restricted to catecholaminergic neurons by means of a region-specific tyrosine hydroxylase promoter, with dosed GDNF expression having trophic effects in dopaminergic neurons (Xia et al. 2008). Most applications of liposomes in the treatment of PD are symptomatic, with efforts devoted to developing optimal pharmacokinetics of l-DOPA and analogues (Spuch and Navarro 2011; Di Stefano et al. 2006). A number of liposomes have been developed as putative AD therapeutics, with a common approach being the use of the vesicle to target (via specific membrane lipids, antibodies or curcumin) and sequester potentially toxic extracellular Aβ (Bereczki et al. 2011; Canovi et al. 2011; Gobbi et al. 2010; Taylor et al. 2011; Mourtas et al. 2011). Intranasal delivery of liposomal formulations of rivastigmine, an acetyl cholinesterase inhibitor used in AD treatment, have been found to exhibit a longer in vivo half-life and effect higher drug concentrations in the brain than the free drug (Mutlu et al. 2011; Arumugam et al. 2008). Furthermore, liposomes have been employed as vaccination vectors against protein misfolding diseases such as AD, by delivering short peptides mimicking pathological epitopes of Aβ or Tau with the intent of eliciting a robust and specific antibody response against the toxic form of the peptide and a subsequent clinical improvement in disease models (Hickman et al. 2011; Muhs et al. 2007; Nicolau et al. 2002; Theunis et al. 2013).
Polymersomes differ from liposomes in the nature of their membrane composition, with the coblock polymers of polymersomes yielding a thicker, more robust membrane (Christian et al. 2009). Accordingly, the membranes of polymersomes are generally considered to be more amenable to modification and functionalization, with many more examples of vesicles with triggered-release mechanisms. This commonly takes the form of a membrane, which degrades in the acidic environment of the endosome, ensuring that large quantities of the polymersome’s therapeutic cargo are delivered into the cytosol (LoPresti et al. 2009). Other vesicles are engineered to release their payload under external stimuli such as UV irradiation, elevated temperature, or appropriate redox conditions (Rijcken et al. 2007; Lee and Feijen 2012). Polymersomes are commonly PEGylated to enhance pharmacokinetics and circulation half-lives, as per their liposomal counterparts, and active targeting is accomplished through the incorporation of antibodies, peptides and small molecule ligands into the external membrane (Christian et al. 2009; Lee and Feijen 2012). One novel example of active targeting involves the incorporation of polyguanylic acid, thus targeting the polymersomes to the macrophage scavenger receptor A1, upregulated in activated tissue macrophages (Broz et al. 2005). Also like liposomes, polymersomes have been most extensively investigated as a means to deliver anti-cancer chemotherapeutics; doxorubicin is once again the archetypical cargo (Waterhouse et al. 2001), however the thick hydrophobic membranes of polymersomes facilitates co-delivery of a more lipophilic drug such as paclitaxel as well (Ahmed et al. 2006). Other putative payloads include genes, RNAi agents, and proteins/enzymes that, with the optimal polymersome composition, remarkably maintain their structure and activity when incorporated into the vesicle membrane or when encapsulated within (Christian et al. 2009). The relatively recent development of polymersomes means that few thoroughly proven examples of their therapeutic utility are available. Nevertheless, they show promise in preliminary applications, including systems of relevance to age-related disease. Polymersomes functionalised with the PR_b peptide, a fibronectin mimetic targeting α5β1 integrin, were developed to deliver siRNA down regulating the Orai3 calcium channel protein in breast cancer cells, inducing cell death (Pangburn et al. 2012). Likewise, incorporation of hyaluronic acid into the external membrane of polymersomes has been found to enhance their delivery to CD44-overexpressing breast cancer cells (Upadhyay et al. 2010), with potential applications in other CD44-overexpressing cancers such as glioma (Knupfer et al. 1999). Transferrin and lactoferrin have also been employed in targeting doxorubicin-laden polymer-somes towards glioma, in vitro and in vivo (Pang et al. 2011; Pang et al. 2010). Lactoferrin-bearing polymersomes were also found to cross the BBB to deliver the neuroprotective peptide S14G-humanin to the neurons of Aβ-treated rats, with a protective effect (Yu et al. 2012). Similarly, polymersomes modified with the TfR antibody OX26 and loaded with NC-1900, a vasopressin fragment analogue known to improve spatial memory impairment, crossed the BBB and accumulated in the brain of scopolamine-lesioned rats, which subsequently performed better in the Morris water maze test (Pang et al. 2008). Studies with neurotensin-modified polyplexes have demonstrated that neurotrophic genes such as GDNF can be delivered with high specificity to neurotensin receptor 1-expressing dopaminergic neurons, a potential therapeutic avenue in PD (Martinez-Fong et al. 2012). More generally, polymersomes functionalised with imaging moieties are also being investigated in the role of diagnostic probes, through the delivery of fluorophores or magnetic resonance imaging (MRI) contrast agents (LoPresti et al. 2009; Levine et al. 2008; Pourtau et al. 2013; Chiang et al. 2013).
Micelles, monolayered spherical arrangements of lipids or (more typically in therapeutic applications) amphiphilic block copolymers, typically with a hydro-phobic core, have seen therapeutic applications in the delivery of poorly water-soluble drugs (Xu et al. 2013). For instance, PEGylated polylactide-based micelles loaded with the hydrophobic drug paclitaxel (Kim et al. 2004) are approved for clinical use in the treatment of breast, lung, and ovarian cancers in South Korea (and undergoing Phase III trials elsewhere) as Genexol-PM. Many other micelle formulations, incorporating a diversity of different chemotherapeutics, are undergoing clinical trials for treating a variety of cancers (Deng et al. 2012). Combinatorial approaches are also being investigated; for instance, micelles loaded with a combined payload of doxorubicin and lapatinib yielded an enhanced doxorubicin uptake in drug-resistant breast cancer cells in vitro, and reduced tumor growth relative to doxorubicin monotherapy in vivo (Wang et al. 2014a). Micelles (polyplexes) have also been employed as nucleic acid delivery agents, typically through incorporation of cationic polymers by which to complex the polyanionic biopolymers (Liu et al. 2013) (Jeong et al. 2011). MRI-active micelles have been used to deliver plasmid DNA into the brain via a compromised BBB after intranasal delivery in mice with traumatic brain injury (Das et al. 2014), while micelles functionalised with the Tat cell-penetrating peptide were able to deliver siRNA to the (intact) brain via the same administration route (Kanazawa et al. 2013). As with the vesicular drug delivery platforms can be engineered with stimulus-responsive drug-release mechanisms (Ganta et al. 2008; Liu et al. 2013), and can be actively targeted by appending and appropriate targeting moiety (e.g. an antibody) to the external polymers (Sawant et al. 2012). BBB-crossing micelles have been developed by targeting nicotine acetylcholine receptors on the capillary endothelium of the brain using the peptide CDX (derived from snake neurotoxin candoxin). These vehicles efficiently delivered paclitaxel into the brain and inhibited intracranial glioblastoma growth in mouse models. Co-delivery of the tumour necrosis-factor related apoptosis-inducing ligand gene enhances the anti-glioblastoma effect even further (Zhan et al. 2012). Haney et al. report on a novel alternative to targeting moieties wherein they macrophages to deliver nanozymes (micelles incorporating the redox enzyme catalase) across an artificial BBB and into microvesicle endothelial cells, neurons, and astrocytes. The antioxidant properties of the nanozymes resulted in efficient reactive oxygen species decomposition, implying potentially useful therapeutic applications in diseases with a neuroinflammatory component, such as AD or PD (Haney et al. 2011).
Various dendritic and polymeric nanoparticle preparations have also been used to deliver chemotherapeutics, nucleic acids, and proteins. Dendrimers are unimolecular, highly branched spherical macromolecules that mimic micelles in their external topology, but also with a liposome-like interior voids (Ganta et al. 2008; Esfand and Tomalia 2001; El Kazzouli et al. 2012). The hydrophilic exterior can be extensively functionalised for enhanced bioavailability and targeting (Kesharwani et al. 2014; Zhu and Shi 2013), a property extensively exploited for tumour-specific delivery of chemotherapeutics (Agarwal et al. 2008). They are also potent siRNA and gene-delivery vectors (Wu et al. 2013). Sialic acid-functionalised dendrimers have been found to mimic cell surface sialic acid clusters, mitigating Aβ-induced neurotoxicity (Patel et al. 2007). Tang et al. describe a dendrimer comprised of multiple hydrolysable l-DOPA units, enhanced in stability and solubility relative to free l-DOPA while facilitating slow release (Tang et al. 2006). This represents a potentially useful prodrug for the treatment of dopamine deficits seen in PD. Dendrimers have also been investigated as potential agents by which to directly destabilise the neurotoxic protein aggregates associated with neurodegenerative disease (Heegaard et al. 2007).
Biodegradable polymeric nanoparticles, commonly based on poly(lactic-co-glycolic acid) (PLGA), have been used to deliver anti-cancer drugs, proteins, genes and RNAi (Danhier et al. 2012). They have also shown promise in treating neurodegenerative disorders (Gao et al. 2013). For instance, PLGA nanoparticles functionalised with the BBB-penetrating peptide TGN delivered the neuroprotective peptide NAP into the brains of mice with model AD, eliciting improved spatial learning and acetylcholinesterase/cholinacetyltransferase activity (Li et al. 2013b). Wheat germ agglutinin-functionalised PLGA nanoparticles enhanced the brain delivery of encapsulated vasoactive intestinal peptide relative to unfunctionalised nanoparticles, and this was again reflected in improved spatial memory and acetylcholinesterase activity in dementia mice (Gao et al. 2007). Lactoferrin-modified polymeric nanoparticles have been used to enhance the delivery of the cytoprotectant urocortin across the BBB of PD-model rats when administered intravenously (Hu et al. 2009; Hu et al. 2011). Odorranalectin has been employed in intranasal administrations for a similar increase in brain-delivery efficiency of urocortin-loaded nanoparticles, alleviating the loss of dopaminergic cells in PD-model rats (Wen et al. 2011).
In recent years, nanostructures that are constructed from nucleic acids have become an increasingly popular prospective therapeutic platform. Such constructs benefit from the fact that both the therapeutic and targeting moieties – siRNA and aptamers, for instance – can be constructed from the same material as the platform without requiring additional synthetic and conjugation steps. RNA nanotechnology (Guo 2010) has demonstrated considerable potential, with one recent approach involving packaging RNA (pRNA) nanostructures, discrete and stable stem-loop structures derived from the DNA-packaging motor of the phi29 bacteriophage (Shu et al. 2014). The pRNA molecule is readily modified to include therapeutic, diagnostic, and targeting moieties, and can be engineered at the supramolecular level to generate multi-meric species via loop–loop interactions (Shu et al. 2013a) or multi-armed junctions (Haque et al. 2012; Shu et al. 2011). Each of the helical arms of these structures can be designed to function as siRNAs, miRNAs, aptamers, or ribozymes, or can be function-alised to append small molecule drugs or ligands, fluorophores, peptides, or additional nanostructures or nanoparticles (Shu et al. 2013b). The crux of the pRNA nanostructure demonstrates remarkable chemical and thermodynamic stability, low immunogenicity and toxicity, and excellent in vivo halflife and biodistribution properties. pRNA nanotechnology has shown considerable promise in laboratory studies, largely in tumour-specific delivery of siRNA. For instance, systemically-delivered folate-functionalised pRNA three-way junctions have been shown to accumulate in folate receptor-overexpressing tumour xenografts in mice (Shu et al. 2011). Similarly, DNA-based self-assembled architectures have also proven highly functionalizable and effective platforms for the delivery of diagnostics and therapeutics (drugs and siRNA/antisense oligonucleotides) in vitro and in vivo (Bhatia et al. 2011; Keum et al. 2011; Walsh et al. 2011; Roh et al. 2011; Lee et al. 2012a; Zhu et al. 2013).
On a larger scale, RNAi microsponges are polymers of repeating hairpin RNAs that self-assemble into pleated sheets, which themselves arrange into a spherical formation. The approximately half-million RNA hairpins are cleavable into siRNA moieties by the RNAi machinery after cellular uptake (with the aid of a polycationic escort), making for potentially potent gene silencing (Lee et al. 2012b). Spherical nucleic acids (SNAs) are densely packed, highly oriented oligonucleotides covalently attached to an inorganic (usually gold) nanoparticle core. They are highly stable and nuclease resistant, capable of autonomous transfection, and capable of strong and persistent gene silencing with minimal immune response or off-target effects (Choi et al. 2013; Cutler et al. 2012). SNAs bearing siRNA targeting the oncoprotein Bcl2Like12, overexpressed in glioblastoma multiforme (GBM), effectively knocked down Bcl2L12 mRNA and protein in GBM cells and, when systemically administered to GBM xenografted mice, increased intratumoral apoptosis and decreased tumor burden (Jensen et al. 2013). As described earlier, SNAs have been delivered via natural exosomes. Exosomal inclusion of SNAs appears to greatly enhance their functional effect: SNAs comprised of anti-miR-21 oligonucleotides were endocytosed into PC-3 prostate cancer cells and subsequently naturally sorted into exosomes. The secreted exosomes, when reintroduced into the same cell type, knocked down the miR-21 oncomiR with an efficacy vastly greater than free SNAs (Alhasan et al. 2014).
While RNA-based therapeutics, particularly those acting via the RNAi pathway, possess great potential, specific and efficient cell delivery to target cells represents a major hurdle to widespread clinical applications. These obstacles are particularly large in the CNS, where the blood–brain and blood–spinal cord barriers obstruct the efficacious uptake of many conventional therapeutics. Encouraging results are being obtained using cell-specific and/or BBB-crossing functionalities, be they nucleic acid aptamers, cell-penetrating peptides, or small molecule receptor-binding ligands, however EVs may be a particularly convenient means to deliver RNA-based therapeutics to their site of action. As described previously, EVs often possess an inherent recipient cell specificity when trafficked in vivo which can be exploited or enhanced for the delivery of therapeutics. Moreover, as delivery vehicles they possess very low immunogenicity, particularly if derived from autologous cells. These aspects are well-illustrated by the DC-derived EV-mediated delivery of siRNA across the BBB and into neural cells, detailed earlier in this review. When administered systemically, these EVs, further functionalised with a brain-penetrating RVG peptide, demonstrated selective and effective delivery of siRNA into the brain (Alvarez-Erviti et al. 2011c). Presently, the in vitro and in vivo delivery of RNAi agents (or other nucleic acids) commonly relies upon generic liposome preparations, the transfection efficacy and toxicity of which varies largely between recipient cell types. It is envisioned that EVs might serve as a means to greatly enhance the efficacy and safety of such transfections, particularly with advances in the ease of EV isolation, purification and characterisation. Thus, EVs represent a potential high-specificity, low-immunogenicity option for the targeted delivery of multifunctional RNA therapeutics.
Conclusions
Our emergent understanding of the diverse and significant roles played by EVs hints at a powerful new therapeutic avenue to exploit in the treatment of disease, including those typically associated with the aging process. This is perhaps most evident in age-related neurodegenerative disorders as EVs tantalise researchers with properties conducive to surmounting the barriers plaguing the generally undruggable ailments of the brain. While advances in bioengineering lead us towards the ability to modulate and adapt naturally occurring EVs, directing their packaging with therapeutic molecules and directing them towards specific recipient cells, concurrent advances in nanoengineering furthers our ability to emulate these properties in synthetic drug-delivery vectors.
Nevertheless, an efficacious exploitation of the EV system will require greater understanding of the intricate in vivo intercellular communication network that extends well beyond the unidirectional processes studied in most in vitro systems. This is of particular significance in multi-factorial, complex conditions such as cellular senescence and organismal aging, especially in immune-specialised environments such as the CNS. Moreover, elucidation of the natural processes of biogenesis, packaging and trafficking will serve to strengthen our efforts at effecting targeted therapy via synthetic routes. Thus, while studies into the nature of EVs are still in their infancy, the significance of these agents of extracellular communication in matters physiological and pathological is abundantly clear.
Acknowledgments
This work was supported by grants from the National Multiple Sclerosis Society (NMSS; RG-4001-A1 to SP), the Italian Multiple Sclerosis Foundation (FISM; RG 2010/R/31 to SP), the Italian Ministry of Health (GR08/7 to SP) the European Research Council (ERC) 2010-StG (RG 260511-SEM_SEM to SP), the European Community (EC) 7th Framework Program (FP7/2007-2013; RG 280772-iONE to SP), The Evelyn Trust (RG 69865 to SP). NI was supported by a FEBS long-term fellowship and BH was supported by China Scholarship Council (No. 201306320024).
Contributor Information
J. A. Smith, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK
T. Leonardi, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK; European Molecular Biology Laboratory, European Bioinformatics Institute (EMBL-EBI), Wellcome Trust Genome Campus, Hinxton, Cambridge CB10 1SD, UK
B. Huang, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK
N. Iraci, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK
B. Vega, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK
S. Pluchino, Department of Clinical Neurosciences, John van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK; Wellcome Trust-Medical Research Council Stem Cell Institute, Cambridge, UK
References
- Agarwal A, Asthana A, Gupta U, Jain NK. Tumour and dendrimers: a review on drug delivery aspects. J Pharm Pharmacol. 2008;60(6):671–688. doi: 10.1211/jpp.60.6.0001. doi:10.1211/jpp.60.6.0001. [DOI] [PubMed] [Google Scholar]
- Aguzzi A, Rajendran L. The transcellular spread of cytosolic amyloids, prions, and prionoids. Neuron. 2009;64(6):783–790. doi: 10.1016/j.neuron.2009.12.016. doi:10.1016/j.neuron.2009.12.016. [DOI] [PubMed] [Google Scholar]
- Ahmed F, Pakunlu RI, Brannan A, Bates F, Minko T, Discher DE. Biodegradable polymersomes loaded with both paclitaxel and doxorubicin permeate and shrink tumors, inducing apoptosis in proportion to accumulated drug. J Control Release. 2006;116(2):150–158. doi: 10.1016/j.jconrel.2006.07.012. doi:10.1016/j.jconrel.2006.07.012. [DOI] [PubMed] [Google Scholar]
- Akbarzadeh A, Rezaei-Sadabady R, Davaran S, Joo SW, Zarghami N, Hanifehpour Y, Samiei M, Kouhi M, Nejati-Koshki K. Liposome: classification, preparation, and applications. Nanoscale Res Lett. 2013;8(1):102. doi: 10.1186/1556-276X-8-102. doi:10.1186/1556-276X-8-102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akers JC, Gonda D, Kim R, Carter BS, Chen CC. Biogenesis of extracellular vesicles (EV): exosomes, microvesicles, retrovirus-like vesicles, and apoptotic bodies. J Neurooncol. 2013;113(1):1–11. doi: 10.1007/s11060-013-1084-8. doi:10.1007/s11060-013-1084-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alais S, Simoes S, Baas D, Lehmann S, Raposo G, Darlix JL, Leblanc P. Mouse neuroblastoma cells release prion infectivity associated with exosomal vesicles. Biol Cell. 2008;100(10):603–615. doi: 10.1042/BC20080025. doi:10.1042/bc20080025. [DOI] [PubMed] [Google Scholar]
- Alegre-Abarrategui J, Christian H, Lufino MM, Mutihac R, Venda LL, Ansorge O, Wade-Martins R. LRRK2 regulates autophagic activity and localizes to specific membrane microdomains in a novel human genomic reporter cellular model. Hum Mol Genet. 2009;18(21):4022–4034. doi: 10.1093/hmg/ddp346. doi:10.1093/hmg/ddp346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alhasan AH, Patel PC, Choi CH, Mirkin CA. Exosome encased spherical nucleic acid gold nanoparticle conjugates as potent microRNA regulation agents. Small. 2014;10(1):186–192. doi: 10.1002/smll.201302143. doi:10.1002/smll.201302143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allen TM, Cullis PR. Liposomal drug delivery systems: from concept to clinical applications. Adv Drug Deliv Rev. 2013;65(1):36–48. doi: 10.1016/j.addr.2012.09.037. doi:10.1016/j.addr.2012.09.037. [DOI] [PubMed] [Google Scholar]
- Al-Nedawi K, Meehan B, Micallef J, Lhotak V, May L, Guha A, Rak J. Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles derived from tumour cells. Nat Cell Biol. 2008;10(5):619–624. doi: 10.1038/ncb1725. doi:10.1038/ncb1725. [DOI] [PubMed] [Google Scholar]
- Alvarez-Erviti L, Seow Y, Schapira AH, Gardiner C, Sargent IL, Wood MJ, Cooper JM. Lysosomal dysfunction increases exosome-mediated alpha-synuclein release and transmission. Neurobiol Dis. 2011a;42(3):360–367. doi: 10.1016/j.nbd.2011.01.029. doi:10.1016/j.nbd.2011.01.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alvarez-Erviti L, Seow Y, Yin H, Betts C, Lakhal S, Wood MJ. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat Biotechnol. 2011b;29(4):341–345. doi: 10.1038/nbt.1807. doi:10.1038/nbt.1807. [DOI] [PubMed] [Google Scholar]
- Alvarez-Erviti L, Seow Y, Yin H, Betts C, Lakhal S, Wood MJA. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat Biotechnol. 2011c;29(4):341–345. doi: 10.1038/nbt.1807. [DOI] [PubMed] [Google Scholar]
- Andreola G, Rivoltini L, Castelli C, Huber V, Perego P, Deho P, Squarcina P, Accornero P, Lozupone F, Lugini L, Stringaro A, Molinari A, Arancia G, Gentile M, Parmiani G, Fais S. Induction of lymphocyte apoptosis by tumor cell secretion of FasL-bearing microvesicles. J Exp Med. 2002;195(10):1303–1316. doi: 10.1084/jem.20011624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Antonyak MA, Li B, Boroughs LK, Johnson JL, Druso JE, Bryant KL, Holowka DA, Cerione RA. Cancer cell-derived microvesicles induce transformation by transferring tissue transglutaminase and fibronectin to recipient cells. Proc Natl Acad Sci USA. 2011;108(12):4852–4857. doi: 10.1073/pnas.1017667108. doi:10.1073/pnas.1017667108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arnold PY, Mannie MD. Vesicles bearing MHC class II molecules mediate transfer of antigen from antigen-presenting cells to CD4 + T cells. Eur J Immunol. 1999;29(4):1363–1373. doi: 10.1002/(SICI)1521-4141(199904)29:04<1363::AID-IMMU1363>3.0.CO;2-0. [DOI] [PubMed] [Google Scholar]
- Arumugam K, Subramanian GS, Mallayasamy SR, Averineni RK, Reddy MS, Udupa N. A study of rivastigmine liposomes for delivery into the brain through intranasal route. Acta Pharm. 2008;58(3):287–297. doi: 10.2478/v10007-008-0014-3. doi:10.2478/v10007-008-0014-3. [DOI] [PubMed] [Google Scholar]
- Aushev VN, Zborovskaya IB, Laktionov KK, Girard N, Cros MP, Herceg Z, Krutovskikh V. Comparisons of microRNA patterns in plasma before and after tumor removal reveal new biomarkers of lung squamous cell carcinoma. PLoS One. 2013;8(10):10. doi: 10.1371/journal.pone.0078649. doi:10.1371/journal.pone.0078649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Azmi AS, Bao B, Sarkar FH. Exosomes in cancer development, metastasis, and drug resistance: a comprehensive review. Cancer Metastasis Rev. 2013;32(3-4):623–642. doi: 10.1007/s10555-013-9441-9. doi:10.1007/s10555-013-9441-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baietti MF, Zhang Z, Mortier E, Melchior A, Degeest G, Geeraerts A, Ivarsson Y, Depoortere F, Coomans C, Vermeiren E, Zimmermann P, David G. Syndecan–syntenin–ALIX regulates the biogenesis of exosomes. Nat Cell Biol. 2012;14(7):677–685. doi: 10.1038/ncb2502. doi:10.1038/ncb2502. [DOI] [PubMed] [Google Scholar]
- Bakhti M, Winter C, Simons M. Inhibition of myelin membrane sheath formation by oligodendrocyte-derived exosome-like vesicles. J Biol Chem. 2011;286(1):787–796. doi: 10.1074/jbc.M110.190009. doi:10.1074/jbc.M110.190009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bala S, Petrasek J, Mundkur S, Catalano D, Levin I, Ward J, Alao H, Kodys K, Szabo G. Circulating microRNAs in exosomes indicate hepatocyte injury and inflammation in alcoholic, drug-induced, and inflammatory liver diseases. Hepatology. 2012;56(5):1946–1957. doi: 10.1002/hep.25873. doi:10.1002/hep.25873. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Banigan MG, Kao PF, Kozubek JA, Winslow AR, Medina J, Costa J, Schmitt A, Schneider A, Cabral H, Cagsal-Getkin O, Vanderburg CR, Delalle I. Differential expression of exosomal microRNAs in prefrontal cortices of schizophrenia and bipolar disorder patients. PLoS One. 2013;8(1):e48814. doi: 10.1371/journal.pone.0048814. doi:10.1371/journal.pone.0048814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bard MP, Hegmans JP, Hemmes A, Luider TM, Willemsen R, Severijnen LA, van Meerbeeck JP, Burgers SA, Hoogsteden HC, Lambrecht BN. Proteomic analysis of exosomes isolated from human malignant pleural effusions. Am J Respir Cell Mol Biol. 2004;31(1):114–121. doi: 10.1165/rcmb.2003-0238OC. doi:10.1165/rcmb.2003-0238OC. [DOI] [PubMed] [Google Scholar]
- Barenholz Y. Doxil(R)—the first FDA-approved nano-drug: lessons learned. J Control Release. 2012;160(2):117–134. doi: 10.1016/j.jconrel.2012.03.020. doi:10.1016/j.jconrel.2012.03.020. [DOI] [PubMed] [Google Scholar]
- Batagov AO, Kuznetsov VA, Kurochkin IV. Identification of nucleotide patterns enriched in secreted RNAs as putative cis-acting elements targeting them to exosome nano-vesicles. BMC Genom. 2011;12(Suppl 3):S18. doi: 10.1186/1471-2164-12-S3-S18. doi:10.1186/1471-2164-12-S3-S18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Batist G, Sawyer M, Gabrail N, Christiansen N, Marshall JL, Spigel DR, Louie A. A multicenter, phase II study of CPX-1 liposome injection in patients (pts) with advanced colorectal cancer (CRC) J Clin Oncol. 2008;26(15):4108. [Google Scholar]
- Bedford P, Garner K, Knight SC. MHC class II molecules transferred between allogeneic dendritic cells stimulate primary mixed leukocyte reactions. Int Immunol. 1999;11(11):1739–1744. doi: 10.1093/intimm/11.11.1739. doi:10.1093/intimm/11.11.1739. [DOI] [PubMed] [Google Scholar]
- Bellingham SA, Coleman BM, Hill AF. Small RNA deep sequencing reveals a distinct miRNA signature released in exosomes from prion-infected neuronal cells. Nucleic Acids Res. 2012a;40(21):10937–10949. doi: 10.1093/nar/gks832. doi:10.1093/nar/gks832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bellingham SA, Guo BB, Coleman BM, Hill AF. Exosomes: vehicles for the transfer of toxic proteins associated with neurodegenerative diseases? Front Physiol. 2012b;3:124. doi: 10.3389/fphys.2012.00124. doi:10.3389/fphys.2012.00124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Belting M, Wittrup A. Nanotubes, exosomes, and nucleic acid-binding peptides provide novel mechanisms of inter-cellular communication in eukaryotic cells: implications in health and disease. J Cell Biol. 2008;183(7):1187–1191. doi: 10.1083/jcb.200810038. doi:10.1083/jcb.200810038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bereczki E, Re F, Masserini ME, Winblad B, Pei JJ. Liposomes functionalized with acidic lipids rescue Abeta-induced toxicity in murine neuroblastoma cells. Nano-medicine. 2011;7(5):560–571. doi: 10.1016/j.nano.2011.05.009. doi:10.1016/j.nano.2011.05.009. [DOI] [PubMed] [Google Scholar]
- Bevers EM, Comfurius P, Dekkers DW, Zwaal RF. Lipid translocation across the plasma membrane of mammalian cells. Biochim Biophys Acta. 1999;1439(3):317–330. doi: 10.1016/s1388-1981(99)00110-9. doi:10.1016/S1388-1981(99)00110-9. [DOI] [PubMed] [Google Scholar]
- Bhatia D, Surana S, Chakraborty S, Koushika SP, Krishnan Y. A synthetic icosahedral DNA-based host–cargo complex for functional in vivo imaging. Nat Commun. 2011;2:339. doi: 10.1038/ncomms1337. doi:10.1038/ncomms1337. [DOI] [PubMed] [Google Scholar]
- Bianco F, Pravettoni E, Colombo A, Schenk U, Moller T, Matteoli M, Verderio C. Astrocyte-derived ATP induces vesicle shedding and IL-1 beta release from microglia. J Immunol. 2005;174(11):7268–7277. doi: 10.4049/jimmunol.174.11.7268. [DOI] [PubMed] [Google Scholar]
- Bianco F, Perrotta C, Novellino L, Francolini M, Riganti L, Menna E, Saglietti L, Schuchman EH, Furlan R, Clementi E, Matteoli M, Verderio C. Acid sphingomyelinase activity triggers microparticle release from glial cells. EMBO J. 2009;28(8):1043–1054. doi: 10.1038/emboj.2009.45. doi:10.1038/emboj.2009.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bieda K, Hoffmann A, Boller K. Phenotypic heterogeneity of human endogenous retrovirus particles produced by teratocarcinoma cell lines. J Gen Virol. 2001;82(Pt 3):591–596. doi: 10.1099/0022-1317-82-3-591. [DOI] [PubMed] [Google Scholar]
- Blume G, Cevc G, Crommelin MD, Bakker-Woudenberg IA, Kluft C, Storm G. Specific targeting with poly(ethylene glycol)-modified liposomes: coupling of homing devices to the ends of the polymeric chains combines effective target binding with long circulation times. Biochim Biophys Acta. 1993;1149(1):180–184. doi: 10.1016/0005-2736(93)90039-3. doi:10.1016/0005-2736(93)90039-3. [DOI] [PubMed] [Google Scholar]
- Boado RJ. Blood–brain barrier transport of non-viral gene and RNAi therapeutics. Pharm Res. 2007;24(9):1772–1787. doi: 10.1007/s11095-007-9321-5. doi:10.1007/s11095-007-9321-5. [DOI] [PubMed] [Google Scholar]
- Bobrie A, Colombo M, Raposo G, Thery C. Exosome secretion: molecular mechanisms and roles in immune responses. Traffic. 2011;12(12):1659–1668. doi: 10.1111/j.1600-0854.2011.01225.x. doi:10.1111/j.1600-0854.2011.01225.x. [DOI] [PubMed] [Google Scholar]
- Boller K, Konig H, Sauter M, Mueller-Lantzsch N, Lower R, Lower J, Kurth R. Evidence that HERV-K is the endogenous retrovirus sequence that codes for the human teratocarcinoma-derived retrovirus HTDV. Virology. 1993;196(1):349–353. doi: 10.1006/viro.1993.1487. doi:10.1006/viro.1993.1487. [DOI] [PubMed] [Google Scholar]
- Bolukbasi MF, Mizrak A, Ozdener GB, Madlener S, Strobel T, Erkan EP, Fan JB, Breakefield XO, Saydam O. miR-1289 and “Zipcode”-like sequence enrich mRNAs in microvesicles. Mol Ther Nucleic Acids. 2012;1:e10. doi: 10.1038/mtna.2011.2. doi:10.1038/mtna.2011.2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bradbury PA, Shepherd FA. Immunotherapy for lung cancer. J Thorac Oncol. 2008;3(6 Suppl 2):S164–S170. doi: 10.1097/JTO.0b013e318174e9a7. doi:10.1097/JTO.0b013e318174e9a7. [DOI] [PubMed] [Google Scholar]
- Brase JC, Johannes M, Schlomm T, Falth M, Haese A, Steuber T, Beissbarth T, Kuner R, Sultmann H. Circulating miRNAs are correlated with tumor progression in prostate cancer. Int J Cancer. 2011;128(3):608–616. doi: 10.1002/ijc.25376. doi:10.1002/ijc.25376. [DOI] [PubMed] [Google Scholar]
- Bronson DL, Fraley EE, Fogh J, Kalter SS. Induction of retrovirus particles in human testicular tumor (Tera-1) cell cultures: an electron microscopic study. J Natl Cancer Inst. 1979;63(2):337–339. [PubMed] [Google Scholar]
- Brouwers JF, Aalberts M, Jansen JW, van Niel G, Wauben MH, Stout TA, Helms JB, Stoorvogel W. Distinct lipid compositions of two types of human prostasomes. Proteomics. 2013;13(10-11):1660–1666. doi: 10.1002/pmic.201200348. doi:10.1002/pmic.201200348. [DOI] [PubMed] [Google Scholar]
- Brown K, Mastrianni JA. The prion diseases. J Geriatr Psychiatry Neurol. 2010;23(4):277–298. doi: 10.1177/0891988710383576. doi:10.1177/0891988710383576. [DOI] [PubMed] [Google Scholar]
- Broz P, Benito SM, Saw C, Burger P, Heider H, Pfisterer M, Marsch S, Meier W, Hunziker P. Cell targeting by a generic receptor-targeted polymer nanocontainer platform. J Control Release. 2005;102(2):475–488. doi: 10.1016/j.jconrel.2004.10.014. doi:10.1016/j.jconrel.2004.10.014. [DOI] [PubMed] [Google Scholar]
- Bryant RJ, Pawlowski T, Catto JW, Marsden G, Vessella RL, Rhees B, Kuslich C, Visakorpi T, Hamdy FC. Changes in circulating microRNA levels associated with prostate cancer. Br J Cancer. 2012;106(4):768–774. doi: 10.1038/bjc.2011.595. doi:10.1038/bjc.2011.595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buyens K, De Smedt SC, Braeckmans K, Demeester J, Peeters L, van Grunsven LA, de Mollerat du Jeu X, Sawant R, Torchilin V, Farkasova K, Ogris M, Sanders NN. Liposome based systems for systemic siRNA delivery: stability in blood sets the requirements for optimal carrier design. J Control Release. 2012;158(3):362–370. doi: 10.1016/j.jconrel.2011.10.009. doi:10.1016/j.jconrel.2011.10.009. [DOI] [PubMed] [Google Scholar]
- Camacho L, Guerrero P, Marchetti D. MicroRNA and protein profiling of brain metastasis competent cell-derived exosomes. PLoS One. 2013;8(9):e73790. doi: 10.1371/journal.pone.0073790. doi:10.1371/journal.pone.0073790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Canovi M, Markoutsa E, Lazar AN, Pampalakis G, Clemente C, Re F, Sesana S, Masserini M, Salmona M, Duyckaerts C, Flores O, Gobbi M, Antimisiaris SG. The binding affinity of anti-Abeta1-42 MAb-decorated nanoliposomes to Abeta1-42 peptides in vitro and to amyloid deposits in post-mortem tissue. Biomaterials. 2011;32(23):5489–5497. doi: 10.1016/j.biomaterials.2011.04.020. doi:10.1016/j.biomaterials.2011.04.020. [DOI] [PubMed] [Google Scholar]
- Cantaluppi V, Gatti S, Medica D, Figliolini F, Bruno S, Deregibus MC, Sordi A, Biancone L, Tetta C, Camussi G. Microvesicles derived from endothelial progenitor cells protect the kidney from ischemia-reperfusion injury by microRNA-dependent reprogramming of resident renal cells. Kidney Int. 2012;82(4):412–427. doi: 10.1038/ki.2012.105. doi:10.1038/ki.2012.105. [DOI] [PubMed] [Google Scholar]
- Cao Z, Tong R, Mishra A, Xu W, Wong GC, Cheng J, Lu Y. Reversible cell-specific drug delivery with aptamer-functionalized liposomes. Angew Chem Int Ed. 2009;48(35):6494–6498. doi: 10.1002/anie.200901452. doi:10.1002/anie.200901452. [DOI] [PubMed] [Google Scholar]
- Carayon K, Chaoui K, Ronzier E, Lazar I, Bertrand-Michel J, Roques V, Balor S, Terce F, Lopez A, Salomé L, Joly E. Proteolipidic composition of exosomes changes during reticulocyte maturation. J Biol Chem. 2011;286(39):34426–34439. doi: 10.1074/jbc.M111.257444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ceruti S, Colombo L, Magni G, Vigano F, Boccazzi M, Deli MA, Sperlagh B, Abbracchio MP, Kittel A. Oxygen–glucose deprivation increases the enzymatic activity and the microvesicle-mediated release of ectonucleotidases in the cells composing the blood–brain barrier. Neurochem Int. 2011;59(2):259–271. doi: 10.1016/j.neuint.2011.05.013. doi:10.1016/j.neuint.2011.05.013. [DOI] [PubMed] [Google Scholar]
- Chandrawati R, Caruso F. Biomimetic liposome- and polymersome-based multicompartmentalized assemblies. Langmuir. 2012;28(39):13798–13807. doi: 10.1021/la301958v. doi:10.1021/la301958v. [DOI] [PubMed] [Google Scholar]
- Chiang WH, Huang WC, Chang CW, Shen MY, Shih ZF, Huang YF, Lin SC, Chiu HC. Functionalized polymer-somes with outlayered polyelectrolyte gels for potential tumor-targeted delivery of multimodal therapies and MR imaging. J Control Release. 2013;168(3):280–288. doi: 10.1016/j.jconrel.2013.03.029. doi:10.1016/j.jconrel.2013.03.029. [DOI] [PubMed] [Google Scholar]
- Chironi GN, Boulanger CM, Simon A, Dignat-George F, Freyssinet JM, Tedgui A. Endothelial microparticles in diseases. Cell Tissue Res. 2009;335(1):143–151. doi: 10.1007/s00441-008-0710-9. doi:10.1007/s00441-008-0710-9. [DOI] [PubMed] [Google Scholar]
- Choi CH, Hao L, Narayan SP, Auyeung E, Mirkin CA. Mechanism for the endocytosis of spherical nucleic acid nanoparticle conjugates. Proc Natl Acad Sci USA. 2013;110(19):7625–7630. doi: 10.1073/pnas.1305804110. doi:10.1073/pnas.1305804110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christian DA, Cai S, Bowen DM, Kim Y, Pajerowski JD, Discher DE. Polymersome carriers: from self-assembly to siRNA and protein therapeutics. Eur J Pharm Biopharm. 2009;71(3):463–474. doi: 10.1016/j.ejpb.2008.09.025. doi:10.1016/j.ejpb.2008.09.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christianson HC, Svensson KJ, van Kuppevelt TH, Li JP, Belting M. Cancer cell exosomes depend on cell-surface heparan sulfate proteoglycans for their internalization and functional activity. Proc Natl Acad Sci USA. 2013;110(43):17380–17385. doi: 10.1073/pnas.1304266110. doi:10.1073/pnas.1304266110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clayton A, Turkes A, Navabi H, Mason MD, Tabi Z. Induction of heat shock proteins in B-cell exosomes. J Cell Sci. 2005;118(Pt 16):3631–3638. doi: 10.1242/jcs.02494. doi:10.1242/jcs.02494. [DOI] [PubMed] [Google Scholar]
- Clayton A, Mitchell JP, Court J, Mason MD, Tabi Z. Human tumor-derived exosomes selectively impair lymphocyte responses to interleukin-2. Cancer Res. 2007;67(15):7458–7466. doi: 10.1158/0008-5472.CAN-06-3456. doi:10.1158/0008-5472.CAN-06-3456. [DOI] [PubMed] [Google Scholar]
- Cocucci E, Racchetti G, Meldolesi J. Shedding micro-vesicles: artefacts no more. Trends Cell Biol. 2009;19(2):43–51. doi: 10.1016/j.tcb.2008.11.003. doi:10.1016/j.tcb.2008.11.003. [DOI] [PubMed] [Google Scholar]
- Cogswell JP, Ward J, Taylor IA, Waters M, Shi YL, Cannon B, Kelnar K, Kemppainen J, Brown D, Chen C, Prinjha RK, Richardson JC, Saunders AM, Roses AD, Richards CA. Identification of miRNA changes in Alzheimer’s disease brain and CSF yields putative biomarkers and insights into disease pathways. J Alzheimers Dis. 2008;14(1):27–41. doi: 10.3233/jad-2008-14103. [DOI] [PubMed] [Google Scholar]
- Coleman BM, Hanssen E, Lawson VA, Hill AF. Prion-infected cells regulate the release of exosomes with distinct ultrastructural features. FASEB J. 2012;26(10):4160–4173. doi: 10.1096/fj.11-202077. doi:10.1096/fj.11-202077. [DOI] [PubMed] [Google Scholar]
- Collet G, Grillon C, Nadim M, Kieda C. Trojan horse at cellular level for tumor gene therapies. Gene. 2013;525(2):208–216. doi: 10.1016/j.gene.2013.03.057. doi:10.1016/j.gene.2013.03.057. [DOI] [PubMed] [Google Scholar]
- Cossetti C, Smith JA, Iraci N, Leonardi T, Alfaro-Cervello C, Pluchino S. Extracellular membrane vesicles and immune regulation in the brain. Front Physiol. 2012;3:117. doi: 10.3389/fphys.2012.00117. doi:10.3389/fphys.2012.00117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cutler JI, Auyeung E, Mirkin CA. Spherical nucleic acids. J Am Chem Soc. 2012;134(3):1376–1391. doi: 10.1021/ja209351u. doi:10.1021/ja209351u. [DOI] [PubMed] [Google Scholar]
- Danhier F, Ansorena E, Silva JM, Coco R, Le Breton A, Preat V. PLGA-based nanoparticles: an overview of biomedical applications. J Control Release. 2012;161(2):505–522. doi: 10.1016/j.jconrel.2012.01.043. doi:10.1016/j.jconrel.2012.01.043. [DOI] [PubMed] [Google Scholar]
- Danzer KM, Kranich LR, Ruf WP, Cagsal-Getkin O, Winslow AR, Zhu L, Vanderburg CR, McLean PJ. Exosomal cell-to-cell transmission of alpha synuclein oligomers. Mol Neurodegener. 2012;7:42. doi: 10.1186/1750-1326-7-42. doi:10.1186/1750-1326-7-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Das M, Wang C, Bedi R, Mohapatra SS, Mohapatra S. Magnetic micelles for DNA delivery to rat brains after mild traumatic brain injury. Nanomedicine. 2014 doi: 10.1016/j.nano.2014.01.003. doi:10.1016/j.nano.2014.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Jong OG, Verhaar MC. Cellular stress conditions are reflected in the protein and RNA content of endothelial cell-derived exosomes. J Extracell Vesicles. 2012;1:18396. doi: 10.3402/jev.v1i0.18396. doi:10.3402/jev.v1i0.18396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Del Conde I, Shrimpton CN, Thiagarajan P, Lopez JA. Tissue-factor-bearing microvesicles arise from lipid rafts and fuse with activated platelets to initiate coagulation. Blood. 2005;106(5):1604–1611. doi: 10.1182/blood-2004-03-1095. doi:10.1182/blood-2004-03-1095. [DOI] [PubMed] [Google Scholar]
- Deng C, Jiang Y, Cheng R, Meng F, Zhong Z. Biodegradable polymeric micelles for targeted and controlled anticancer drug delivery: promises, progress and prospects. Nano Today. 2012;7(5):467–480. doi:10.1016/j.nantod.2012.08.005. [Google Scholar]
- Depil S, Roche C, Dussart P, Prin L. Expression of a human endogenous retrovirus, HERV-K, in the blood cells of leukemia patients. Leukemia. 2002;16(2):254–259. doi: 10.1038/sj.leu.2402355. doi:10.1038/sj.leu.2402355. [DOI] [PubMed] [Google Scholar]
- DePinho RA. The age of cancer. Nature. 2000;408(6809):248–254. doi: 10.1038/35041694. doi:10.1038/35041694. [DOI] [PubMed] [Google Scholar]
- Deregibus MC, Cantaluppi V, Calogero R, Lo Iacono M, Tetta C, Biancone L, Bruno S, Bussolati B, Camussi G. Endothelial progenitor cell derived microvesicles activate an angiogenic program in endothelial cells by a horizontal transfer of mRNA. Blood. 2007;110(7):2440–2448. doi: 10.1182/blood-2007-03-078709. doi:10.1182/blood-2007-03-078709. [DOI] [PubMed] [Google Scholar]
- Desplats P, Lee HJ, Bae EJ, Patrick C, Rockenstein E, Crews L, Spencer B, Masliah E, Lee SJ. Inclusion formation and neuronal cell death through neuron-to-neuron transmission of alpha-synuclein. Proc Natl Acad Sci USA. 2009;106(31):13010–13015. doi: 10.1073/pnas.0903691106. doi:10.1073/pnas.0903691106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devadasu VR, Bhardwaj V, Kumar MN. Can controversial nanotechnology promise drug delivery? Chem Rev. 2013;113(3):1686–1735. doi: 10.1021/cr300047q. doi:10.1021/cr300047q. [DOI] [PubMed] [Google Scholar]
- Dewannieux M, Blaise S, Heidmann T. Identification of a functional envelope protein from the HERV-K family of human endogenous retroviruses. J Virol. 2005;79(24):15573–15577. doi: 10.1128/JVI.79.24.15573-15577.2005. doi:10.1128/JVI.79.24.15573-15577.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Di Stefano A, Sozio P, Iannitelli A, Marianecci C, Santucci E, Carafa M. Maleic- and fumaric-diamides of (O,O-diacetyl)-L-dopa-methylester as anti-Parkinson prodrugs in liposomal formulation. J Drug Target. 2006;14(9):652–661. doi: 10.1080/10611860600916636. doi:10.1080/10611860600916636. [DOI] [PubMed] [Google Scholar]
- Dihanich S, Manzoni C. LRRK2: a problem lurking in vesicle trafficking? J Neurosci. 2011;31(27):9787–9788. doi: 10.1523/JNEUROSCI.1976-11.2011. doi:10.1523/JNEUROSCI.1976-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dinkins MB, Dasgupta S, Wang G, Zhu G, Bieberich E. Exosome reduction in vivo is associated with lower amyloid plaque load in the 5XFAD mouse model of Alzheimer’s disease. Neurobiol Aging. 2014;35(8):1792–1800. doi: 10.1016/j.neurobiolaging.2014.02.012. doi:10.1016/j.neurobiolaging.2014.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drummond DC, Noble CO, Guo Z, Hong K, Park JW, Kirpotin DB. Development of a highly active nanoliposomal irinotecan using a novel intraliposomal stabilization strategy. Cancer Res. 2006;66(6):3271–3277. doi: 10.1158/0008-5472.CAN-05-4007. doi:10.1158/0008-5472.CAN-05-4007. [DOI] [PubMed] [Google Scholar]
- Duncan R, Gaspar R. Nanomedicine(s) under the microscope. Mol Pharm. 2011;8(6):2101–2141. doi: 10.1021/mp200394t. doi:10.1021/mp200394t. [DOI] [PubMed] [Google Scholar]
- El Kazzouli S, Mignani S, Bousmina M, Majoral J-P. Dendrimer therapeutics: covalent and ionic attachments. New J Chem. 2012;36(2):227. doi:10.1039/c1nj20459a. [Google Scholar]
- El-Andaloussi S, Lee Y, Lakhal-Littleton S, Li J, Seow Y, Gardiner C, Alvarez-Erviti L, Sargent IL, Wood MJ. Exosome-mediated delivery of siRNA in vitro and in vivo. Nat Protoc. 2012;7(12):2112–2126. doi: 10.1038/nprot.2012.131. doi:10.1038/nprot.2012.131. [DOI] [PubMed] [Google Scholar]
- El-Andaloussi S, Mager I, Breakefield XO, Wood MJ. Extracellular vesicles: biology and emerging therapeutic opportunities. Nat Rev Drug Discov. 2013;12(5):347–357. doi: 10.1038/nrd3978. doi:10.1038/nrd3978. [DOI] [PubMed] [Google Scholar]
- Eldh M, Ekström K, Valadi H, Sjöstrand M, Olsson B. Exosomes communicate protective messages during oxidative stress; possible role of exosomal shuttle RNA. PLoS One. 2010;5(12):e15353. doi: 10.1371/journal.pone.0015353. doi:10.1371/journal.pone.0015353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Emmanouilidou E, Melachroinou K, Roumeliotis T, Garbis SD, Ntzouni M, Margaritis LH, Stefanis L, Vekrellis K. Cell-produced alpha-synuclein is secreted in a calcium-dependent manner by exosomes and impacts neuronal survival. J Neurosci. 2010;30(20):6838–6851. doi: 10.1523/JNEUROSCI.5699-09.2010. doi:10.1523/JNEUROSCI.5699-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Escola JM, Kleijmeer MJ, Stoorvogel W, Griffith JM, Yoshie O, Geuze HJ. Selective enrichment of tetraspan proteins on the internal vesicles of multivesicular endosomes and on exosomes secreted by human B-lymphocytes. J Biol Chem. 1998;273(32):20121–20127. doi: 10.1074/jbc.273.32.20121. [DOI] [PubMed] [Google Scholar]
- Esfand R, Tomalia DA. Poly(amidoamine) (PAMAM) dendrimers: from biomimicry to drug delivery and biomedical applications. Drug Discov Today. 2001;6(8):427–436. doi: 10.1016/s1359-6446(01)01757-3. [DOI] [PubMed] [Google Scholar]
- Fantini M, Gianni L, Santelmo C, Drudi F, Castellani C, Affatato A, Nicolini M, Ravaioli A. Lipoplatin treatment in lung and breast cancer. Chemother Res Pract. 2011;2011:125192. doi: 10.1155/2011/125192. doi:10.1155/2011/125192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fasol U, Frost A, Buchert M, Arends J, Fiedler U, Scharr D, Scheuenpflug J, Mross K. Vascular and pharmaco-kinetic effects of EndoTAG-1 in patients with advanced cancer and liver metastasis. Ann Oncol. 2012;23(4):1030–1036. doi: 10.1093/annonc/mdr300. doi:10.1093/annonc/mdr300. [DOI] [PubMed] [Google Scholar]
- Faure J, Lachenal G, Court M, Hirrlinger J, Chatellard-Causse C, Blot B, Grange J, Schoehn G, Goldberg Y, Boyer V, Kirchhoff F, Raposo G, Garin J, Sadoul R. Exosomes are released by cultured cortical neurones. Mol Cell Neurosci. 2006;31(4):642–648. doi: 10.1016/j.mcn.2005.12.003. doi:10.1016/j.mcn.2005.12.003. [DOI] [PubMed] [Google Scholar]
- Feng Z. p53 regulation of the IGF-1/AKT/mTOR pathways and the endosomal compartment. Cold Spring Harb Perspect Biol. 2010;2(2):a001057. doi: 10.1101/cshperspect.a001057. doi:10.1101/cshperspect.a001057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fevrier B, Vilette D, Archer F, Loew D, Faigle W, Vidal M, Laude H, Raposo G. Cells release prions in association with exosomes. Proc Natl Acad Sci USA. 2004;101(26):9683–9688. doi: 10.1073/pnas.0308413101. doi:10.1073/pnas.0308413101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fitzner D, Schnaars M, van Rossum D. Selective transfer of exosomes from oligodendrocytes to microglia by macropinocytosis. J Cell Sci. 2011;124(3):447–458. doi: 10.1242/jcs.074088. [DOI] [PubMed] [Google Scholar]
- Fruhbeis C, Frohlich D, Kuo WP, Amphornrat J, Thilemann S, Saab AS, Kirchhoff F, Mobius W, Goebbels S, Nave KA, Schneider A, Simons M, Klugmann M, Trotter J, Kramer-Albers EM. Neurotransmitter-triggered transfer of exosomes mediates oligodendrocyte-neuron communication. PLoS Biol. 2013a;11(7):e1001604. doi: 10.1371/journal.pbio.1001604. doi:10.1371/journal.pbio.1001604. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fruhbeis C, Frohlich D, Kuo WP, Kramer-Albers EM. Extracellular vesicles as mediators of neuron-glia communication. Front Cell Neurosci. 2013b;7:182. doi: 10.3389/fncel.2013.00182. doi:10.3389/fncel.2013.00182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gabizon A, Catane R, Uziely B, Kaufman B, Safra T, Cohen R, Martin F, Huang A, Barenholz Y. Prolonged circulation time and enhanced accumulation in malignant exudates of doxorubicin encapsulated in polyethylene-glycol coated liposomes. Cancer Res. 1994;54(4):987–992. [PubMed] [Google Scholar]
- Gabriel K, Ingram A, Austin R, Kapoor A, Tang D, Majeed F, Qureshi T, Al-Nedawi K. Regulation of the tumor suppressor PTEN through exosomes: a diagnostic potential for prostate cancer. PLoS One. 2013;8(7):e70047. doi: 10.1371/journal.pone.0070047. doi:10.1371/journal.pone.0070047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ganta S, Devalapally H, Shahiwala A, Amiji M. A review of stimuli-responsive nanocarriers for drug and gene delivery. J Control Release. 2008;126(3):187–204. doi: 10.1016/j.jconrel.2007.12.017. doi:10.1016/j.jconrel.2007.12.017. [DOI] [PubMed] [Google Scholar]
- Gao X, Wu B, Zhang Q, Chen J, Zhu J, Zhang W, Rong Z, Chen H, Jiang X. Brain delivery of vasoactive intestinal peptide enhanced with the nanoparticles conjugated with wheat germ agglutinin following intranasal administration. J Control Release. 2007;121(3):156–167. doi: 10.1016/j.jconrel.2007.05.026. doi:10.1016/j.jconrel.2007.05.026. [DOI] [PubMed] [Google Scholar]
- Gao H, Pang Z, Jiang X. Targeted delivery of nanotherapeutics for major disorders of the central nervous system. Pharm Res. 2013;30(10):2485–2498. doi: 10.1007/s11095-013-1122-4. doi:10.1007/s11095-013-1122-4. [DOI] [PubMed] [Google Scholar]
- Gastpar R, Gehrmann M, Bausero MA, Asea A, Gross C. Heat shock protein 70 surface-positive tumor exosomes stimulate migratory and cytolytic activity of natural killer cells. Cancer Res. 2005;65(12):5238–5247. doi: 10.1158/0008-5472.CAN-04-3804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ge J, Neofytou E, Lei J, Beygui RE, Zare RN. Protein–polymer hybrid nanoparticles for drug delivery. Small. 2012;8(23):3573–3578. doi: 10.1002/smll.201200889. doi:10.1002/smll.201200889. [DOI] [PubMed] [Google Scholar]
- Gildea JJ, Carlson JM, Schoeffel CD, Carey RM, Felder RA. Urinary exosome miRNome analysis and its applications to salt sensitivity of blood pressure. Clin Biochem. 2013;46(12):1131–1134. doi: 10.1016/j.clinbiochem.2013.05.052. doi:10.1016/j.clinbiochem.2013.05.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gladnikoff M, Shimoni E, Gov NS, Rousso I. Retroviral assembly and budding occur through an actin-driven mechanism. Biophys J. 2009;97(9):2419–2428. doi: 10.1016/j.bpj.2009.08.016. doi:10.1016/j.bpj.2009.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gobbi M, Re F, Canovi M, Beeg M, Gregori M, Sesana S, Sonnino S, Brogioli D, Musicanti C, Gasco P, Salmona M, Masserini ME. Lipid-based nanoparticles with high binding affinity for amyloid-beta1-42 peptide. Biomaterials. 2010;31(25):6519–6529. doi: 10.1016/j.biomaterials.2010.04.044. doi:10.1016/j.biomaterials.2010.04.044. [DOI] [PubMed] [Google Scholar]
- Golan M, Hizi A, Resau JH, Yaal-Hahoshen N, Reichman H, Keydar I, Tsarfaty I. Human endogenous retrovirus (HERV-K) reverse transcriptase as a breast cancer prognostic marker. Neoplasia. 2008;10(6):521–533. doi: 10.1593/neo.07986. doi:10.1593/neo.07986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gould SJ, Raposo G. As we wait: coping with an imperfect nomenclature for extracellular vesicles. J Extra-cell Vesicles. 2013;2 doi: 10.3402/jev.v2i0.20389. doi:10.3402/jev.v2i0.20389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grange C, Tapparo M, Collino F, Vitillo L, Damasco C, Deregibus MC, Tetta C, Bussolati B, Camussi G. Microvesicles released from human renal cancer stem cells stimulate angiogenesis and formation of lung premetastatic niche. Cancer Res. 2011;71(15):5346–5356. doi: 10.1158/0008-5472.CAN-11-0241. doi:10.1158/0008-5472.CAN-11-0241. [DOI] [PubMed] [Google Scholar]
- Gregoriadis G, Ryman BE. Liposomes as carriers of enzymes or drugs: a new approach to the treatment of storage diseases. Biochem J. 1971;124(5):58. doi: 10.1042/bj1240058p. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gross JC, Chaudhary V, Bartscherer K, Boutros M. Active Wnt proteins are secreted on exosomes. Nat Cell Biol. 2012;14(10):1036–1045. doi: 10.1038/ncb2574. doi:10.1038/ncb2574. [DOI] [PubMed] [Google Scholar]
- Guescini M, Genedani S, Stocchi V, Agnati LF. Astrocytes and glioblastoma cells release exosomes carrying mtDNA. J Neural Transm. 2010;117(1):1–4. doi: 10.1007/s00702-009-0288-8. doi:10.1007/s00702-009-0288-8. [DOI] [PubMed] [Google Scholar]
- Guo P. The emerging field of RNA nanotechnology. Nat Nanotechnol. 2010;5(12):833–842. doi: 10.1038/nnano.2010.231. doi:10.1038/nnano.2010.231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gupta S, Knowlton AA. HSP60 trafficking in adult cardiac myocytes: role of the exosomal pathway. Am J Physiol Heart Circ Physiol. 2007;292(6):H3052–H3056. doi: 10.1152/ajpheart.01355.2006. doi:10.1152/ajpheart.01355.2006. [DOI] [PubMed] [Google Scholar]
- Hajrasouliha AR, Jiang G, Lu Q, Lu H, Kaplan HJ, Zhang HG, Shao H. Exosomes from retinal astrocytes contain antiangiogenic components that inhibit laser-induced choroidal neovascularization. J Biol Chem. 2013;288(39):28058–28067. doi: 10.1074/jbc.M113.470765. doi:10.1074/jbc.M113.470765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Halliday GM, Stevens CH. Glia: initiators and progressors of pathology in Parkinson’s disease. Mov Disord. 2011;26(1):6–17. doi: 10.1002/mds.23455. doi:10.1002/mds.23455. [DOI] [PubMed] [Google Scholar]
- Haney MJ, Zhao Y, Li S, Higginbotham SM, Booth SL, Han HY, Vetro JA, Mosley RL, Kabanov AV, Gendelman HE, Batrakova EV. Cell-mediated transfer of catalase nanoparticles from macrophages to brain endothelial, glial and neuronal cells. Nanomedicine. 2011;6(7):1215–1230. doi: 10.2217/nnm.11.32. doi:10.2217/nnm.11.32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hansen C, Angot E, Bergstrom AL, Steiner JA, Pieri L, Paul G, Outeiro TF, Melki R, Kallunki P, Fog K, Li JY, Brundin P. Alpha-synuclein propagates from mouse brain to grafted dopaminergic neurons and seeds aggregation in cultured human cells. J Clin Invest. 2011;121(2):715–725. doi: 10.1172/JCI43366. doi:10.1172/JCI43366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haque F, Shu D, Shu Y, Shlyakhtenko LS, Rychahou PG, Evers BM, Guo P. Ultrastable synergistic tetravalent RNA nanoparticles for targeting to cancers. Nano Today. 2012;7(4):245–257. doi: 10.1016/j.nantod.2012.06.010. doi:10.1016/j.nantod.2012.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hawari FI, Rouhani FN, Cui X, Yu ZX, Buckley C, Kaler M, Levine SJ. Release of full-length 55-kDa TNF receptor 1 in exosome-like vesicles: a mechanism for generation of soluble cytokine receptors. Proc Natl Acad Sci USA. 2004;101(5):1297–1302. doi: 10.1073/pnas.0307981100. doi:10.1073/pnas.0307981100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hedlund M, Stenqvist A-C, Nagaeva O, Kjellberg L, Wulff M, Baranov V, Mincheva-Nilsson L. Human placenta expresses and secretes NKG2D ligands via exosomes that down-modulate the cognate receptor expression: evidence for immunosuppressive function. J Immunol. 2009;183(1):340–351. doi: 10.4049/jimmunol.0803477. doi:10.4049/jimmunol.0803477. [DOI] [PubMed] [Google Scholar]
- Heegaard PM, Boas U, Otzen DE. Dendrimer effects on peptide and protein fibrillation. Macromol Biosci. 2007;7(8):1047–1059. doi: 10.1002/mabi.200700051. doi:10.1002/mabi.200700051. [DOI] [PubMed] [Google Scholar]
- Heijnen HF, Schiel AE, Fijnheer R, Geuze HJ, Sixma JJ. Activated platelets release two types of membrane vesicles: microvesicles by surface shedding and exosomes derived from exocytosis of multivesicular bodies and alpha-granules. Blood. 1999;94(11):3791–3799. [PubMed] [Google Scholar]
- Hemler ME. Tetraspanin proteins mediate cellular penetration, invasion, and fusion events and define a novel type of membrane microdomain. Annu Rev Cell Dev Biol. 2003;19:397–422. doi: 10.1146/annurev.cellbio.19.111301.153609. doi:10.1146/annurev.cellbio.19.111301.153609. [DOI] [PubMed] [Google Scholar]
- Hendriks BS, Klinz SG, Reynolds JG, Espelin CW, Gaddy DF, Wickham TJ. Impact of tumor HER2/ERBB2 expression level on HER2-targeted liposomal doxorubicin-mediated drug delivery: multiple low-affinity interactions lead to a threshold effect. Mol Cancer Ther. 2013;12(9):1816–1828. doi: 10.1158/1535-7163.MCT-13-0180. doi:10.1158/1535-7163.MCT-13-0180. [DOI] [PubMed] [Google Scholar]
- Hergenreider E, Heydt S, Treguer K, Boettger T, Horrevoets AJ, Zeiher AM, Scheffer MP, Frangakis AS, Yin X, Mayr M, Braun T, Urbich C, Boon RA, Dimmeler S. Atheroprotective communication between endothelial cells and smooth muscle cells through miRNAs. Nat Cell Biol. 2012;14(3):249–256. doi: 10.1038/ncb2441. doi:10.1038/ncb2441. [DOI] [PubMed] [Google Scholar]
- Hessvik NP, Sandvig K, Llorente A. Exosomal miRNAs as biomarkers for prostate cancer. Front Genet. 2013;4:36. doi: 10.3389/fgene.2013.00036. doi:10.3389/fgene.2013.00036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hickman DT, Lopez-Deber MP, Ndao DM, Silva AB, Nand D, Pihlgren M, Giriens V, Madani R, St-Pierre A, Karastaneva H, Nagel-Steger L, Willbold D, Riesner D, Nicolau C, Baldus M, Pfeifer A, Muhs A. Sequence-independent control of peptide conformation in liposomal vaccines for targeting protein misfolding diseases. J Biol Chem. 2011;286(16):13966–13976. doi: 10.1074/jbc.M110.186338. doi:10.1074/jbc.M110.186338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hsu C, Morohashi Y, Yoshimura S, Manrique-Hoyos N, Jung S, Lauterbach MA, Bakhti M, Gronborg M, Mobius W, Rhee J, Barr FA, Simons M. Regulation of exosome secretion by Rab35 and its GTPase-activating proteins TBC1D10A-C. J Cell Biol. 2010;189(2):223–232. doi: 10.1083/jcb.200911018. doi:10.1083/jcb.200911018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu K, Li J, Shen Y, Lu W, Gao X, Zhang Q, Jiang X. Lactoferrin-conjugated PEG-PLA nanoparticles with improved brain delivery: in vitro and in vivo evaluations. J Control Release. 2009;134(1):55–61. doi: 10.1016/j.jconrel.2008.10.016. doi:10.1016/j.jconrel.2008.10.016. [DOI] [PubMed] [Google Scholar]
- Hu K, Shi Y, Jiang W, Han J, Huang S, Jiang X. Lactoferrin conjugated PEG–PLGA nanoparticles for brain delivery: preparation, characterization and efficacy in Parkinson’s disease. Int J Pharm. 2011;415(1-2):273–283. doi: 10.1016/j.ijpharm.2011.05.062. doi:10.1016/j.ijpharm.2011.05.062. [DOI] [PubMed] [Google Scholar]
- Hu G, Yao H, Chaudhuri AD, Duan M, Yelamanchili SV, Wen H, Cheney PD, Fox HS, Buch S. Exosome-mediated shuttling of microRNA-29 regulates HIV Tat and morphine-mediated Neuronal dysfunction. Cell Death Dis. 2012;3:e381. doi: 10.1038/cddis.2012.114. doi:10.1038/cddis.2012.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huber V, Fais S, Iero M, Lugini L, Canese P, Squarcina P, Zaccheddu A, Colone M, Arancia G, Gentile M, Seregni E, Valenti R, Ballabio G, Belli F, Leo E, Parmiani G, Rivoltini L. Human colorectal cancer cells induce T-cell death through release of proapoptotic microvesicles: role in immune escape. Gastroenterology. 2005;128(7):1796–1804. doi: 10.1053/j.gastro.2005.03.045. doi:10.1053/j.gastro.2005.03.045. [DOI] [PubMed] [Google Scholar]
- Hugel B, Martinez MC, Kunzelmann C, Freyssinet JM. Membrane microparticles: two sides of the coin. Physiology. 2005;20:22–27. doi: 10.1152/physiol.00029.2004. doi:10.1152/physiol.00029.2004. [DOI] [PubMed] [Google Scholar]
- Hung CW, Chen YC, Hsieh WL, Chiou SH, Kao CL. Ageing and neurodegenerative diseases. Ageing Res Rev. 2010;9(Suppl 1):S36–S46. doi: 10.1016/j.arr.2010.08.006. doi:10.1016/j.arr.2010.08.006. [DOI] [PubMed] [Google Scholar]
- Huttner HB, Janich P, Kohrmann M, Jaszai J, Siebzehnrubl F, Blumcke I, Suttorp M, Gahr M, Kuhnt D, Nimsky C, Krex D, Schackert G, Lowenbruck K, Reichmann H, Juttler E, Hacke W, Schellinger PD, Schwab S, Wilsch-Brauninger M, Marzesco AM, Corbeil D. The stem cell marker prominin-1/CD133 on membrane particles in human cerebrospinal fluid offers novel approaches for studying central nervous system disease. Stem Cells. 2008;26(3):698–705. doi: 10.1634/stemcells.2007-0639. doi:10.1634/stemcells.2007-0639. [DOI] [PubMed] [Google Scholar]
- Iero M, Valenti R, Huber V, Filipazzi P, Parmiani G, Fais S, Rivoltini L. Tumour-released exosomes and their implications in cancer immunity. Cell Death Differ. 2007;15(1):80–88. doi: 10.1038/sj.cdd.4402237. doi:10.1038/sj.cdd.4402237. [DOI] [PubMed] [Google Scholar]
- Jang SC, Kim OY, Yoon CM, Choi DS, Roh TY, Park J, Nilsson J, Lotvall J, Kim YK, Gho YS. Bioinspired exo-some-mimetic nanovesicles for targeted delivery of chemotherapeutics to malignant tumors. ACS Nano. 2013;7(9):7698–7710. doi: 10.1021/nn402232g. doi:10.1021/nn402232g. [DOI] [PubMed] [Google Scholar]
- Janowska-Wieczorek A, Wysoczynski M, Kijowski J, Marquez-Curtis L, Machalinski B, Ratajczak J, Ratajczak MZ. Microvesicles derived from activated platelets induce metastasis and angiogenesis in lung cancer. Int J Cancer. 2005;113(5):752–760. doi: 10.1002/ijc.20657. doi:10.1002/ijc.20657. [DOI] [PubMed] [Google Scholar]
- Jayaraman M, Ansell SM, Mui BL, Tam YK, Chen J, Du X, Butler D, Eltepu L, Matsuda S, Narayanannair JK, Rajeev KG, Hafez IM, Akinc A, Maier MA, Tracy MA, Cullis PR, Madden TD, Manoharan M, Hope MJ. Maximizing the potency of siRNA lipid nanoparticles for hepatic gene silencing in vivo. Angew Chem Int Ed. 2012;51(34):8529–8533. doi: 10.1002/anie.201203263. doi:10.1002/anie.201203263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jensen SA, Day ES, Ko CH, Hurley LA, Luciano JP, Kouri FM, Merkel TJ, Luthi AJ, Patel PC, Cutler JI, Daniel WL, Scott AW, Rotz MW, Meade TJ, Giljohann DA, Mirkin CA, Stegh AH. Spherical nucleic acid nanoparticle conjugates as an RNAi-based therapy for glioblastoma. Sci Transl Med. 2013;5(209):209RA152. doi: 10.1126/scitranslmed.3006839. doi:10.1126/scitranslmed.3006839. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jeong JH, Park TG, Kim SH. Self-assembled and nano-structured siRNA delivery systems. Pharm Res. 2011;28(9):2072–2085. doi: 10.1007/s11095-011-0412-y. doi:10.1007/s11095-011-0412-y. [DOI] [PubMed] [Google Scholar]
- Jo W, Jeong D, Kim J, Cho S, Jang SC, Han C, Kang JY, Gho YS, Park J. Microfluidic fabrication of cell-derived nanovesicles as endogenous RNA carriers. Lab Chip. 2014;14(7):1261–1269. doi: 10.1039/c3lc50993a. doi:10.1039/c3lc50993a. [DOI] [PubMed] [Google Scholar]
- Johnstone RM, Adam M, Hammond JR, Orr L, Turbide C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes) J Biol Chem. 1987;262(19):9412–9420. [PubMed] [Google Scholar]
- Joshi P, Turola E, Ruiz A, Bergami A, Libera DD, Benussi L, Giussani P, Magnani G, Comi G, Legname G, Ghidoni R, Furlan R, Matteoli M, Verderio C. Microglia convert aggregated amyloid-beta into neurotoxic forms through the shedding of microvesicles. Cell Death Differ. 2014;21(4):582–593. doi: 10.1038/cdd.2013.180. doi:10.1038/cdd.2013.180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jung KH, Chu K, Lee ST, Park HK, Bahn JJ, Kim DH, Kim JH, Kim M, Kun Lee S, Roh JK. Circulating endothelial microparticles as a marker of cerebrovascular disease. Ann Neurol. 2009;66(2):191–199. doi: 10.1002/ana.21681. doi:10.1002/ana.21681. [DOI] [PubMed] [Google Scholar]
- Jy W, Minagar A, Jimenez JJ, Sheremata WA, Mauro LM, Horstman LL, Bidot C, Ahn YS. Endothelial microparticles (EMP) bind and activate monocytes: elevated EMP-monocyte conjugates in multiple sclerosis. Front Biosci. 2004;9:3137–3144. doi: 10.2741/1466. doi:10.2741/1466. [DOI] [PubMed] [Google Scholar]
- Kalani A, Tyagi A, Tyagi N. Exosomes: mediators of neurodegeneration, neuroprotection and therapeutics. Mol Neurobiol. 2013;49(1):590–600. doi: 10.1007/s12035-013-8544-1. doi:10.1007/s12035-013-8544-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kalra H, Simpson RJ, Ji H, Aikawa E, Altevogt P, Askenase P, Bond VC, Borràs FE, Breakefield X, Budnik V, Buzas E, Camussi G, Clayton A, Cocucci E, Falcon-Perez JM, Gabrielsson S, Gho YS, Gupta D, Harsha HC, Hendrix A, Hill AF, Inal JM, Jenster G, Krämer-Albers E-M, Lim SK, Llorente A, Lotvall J, Marcilla A, Mincheva-Nilsson L, Nazarenko I, Nieuwland R, Nolte-’t Hoen ENM, Pandey A, Patel T, Piper MG, Pluchino S, Prasad TSK, Rajendran L, Raposo G, Record M, Reid GE, Sánchez-Madrid F, Schiffelers RM, Siljander P, Stensballe A, Stoorvogel W, Taylor D, Théry C, Valadi H, van Balkom BWM, Vázquez J, Vidal M, Wauben MHM, Yáñez-Mó M, Zoeller M, Mathivanan S. Vesiclepedia: a compendium for extracellular vesicles with continuous community annotation. PLoS Biol. 2012;10(12):e1001450. doi: 10.1371/journal.pbio.1001450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanasty R, Dorkin JR, Vegas A, Anderson D. Delivery materials for siRNA therapeutics. Nat Mater. 2013;12(11):967–977. doi: 10.1038/nmat3765. doi:10.1038/nmat3765. [DOI] [PubMed] [Google Scholar]
- Kanazawa T, Akiyama F, Kakizaki S, Takashima Y, Seta Y. Delivery of siRNA to the brain using a combination of nose-to-brain delivery and cell-penetrating peptide-modified nano-micelles. Biomaterials. 2013;34(36):9220–9226. doi: 10.1016/j.biomaterials.2013.08.036. doi:10.1016/j.biomaterials.2013.08.036. [DOI] [PubMed] [Google Scholar]
- Karolina DS, Tavintharan S, Armugam A, Sepramaniam S, Pek SLT, Wong MTK, Lim SC, Sum CF, Jeyaseelan K. Circulating miRNA profiles in patients with metabolic syndrome. J Clin Endocrinol Metab. 2012;97(12):E2271–E2276. doi: 10.1210/jc.2012-1996. doi:10.1210/jc.2012-1996. [DOI] [PubMed] [Google Scholar]
- Katakowski M, Buller B, Zheng X, Lu Y, Rogers T, Osobamiro O, Shu W, Jiang F, Chopp M. Exosomes from marrow stromal cells expressing miR-146b inhibit glioma growth. Cancer Lett. 2013;335(1):201–204. doi: 10.1016/j.canlet.2013.02.019. doi:10.1016/j.canlet.2013.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kesharwani P, Jain K, Jain NK. Dendrimer as nanocarrier for drug delivery. Prog Polym Sci. 2014;39(2):268–307. doi:10.1016/j.progpolymsci.2013.07.005. [Google Scholar]
- Keum JW, Ahn JH, Bermudez H. Design, assembly, and activity of antisense DNA nanostructures. Small. 2011;7(24):3529–3535. doi: 10.1002/smll.201101804. doi:10.1002/smll.201101804. [DOI] [PubMed] [Google Scholar]
- Kim TY, Kim DW, Chung JY, Shin SG, Kim SC, Heo DS, Kim NK, Bang YJ. Phase I and pharmacokinetic study of Genexol-PM, a cremophor-free, polymeric micelle-formulated paclitaxel, in patients with advanced malignancies. Clin Cancer Res. 2004;10(11):3708–3716. doi: 10.1158/1078-0432.CCR-03-0655. doi:10.1158/1078-0432.CCR-03-0655. [DOI] [PubMed] [Google Scholar]
- Knupfer MM, Poppenborg H, Hotfilder M, Kuhnel K, Wolff JE, Domula M. CD44 expression and hyaluronic acid binding of malignant glioma cells. Clin Exp Metastasis. 1999;17(1):71–76. doi: 10.1023/a:1026425519497. doi:10.1023/A:1026425519497. [DOI] [PubMed] [Google Scholar]
- Kobayashi M, Salomon C, Tapia J, Illanes SE, Mitchell MD, Rice GE. Ovarian cancer cell invasiveness is associated with discordant exosomal sequestration of Let-7 miRNA and miR-200. J Transl Med. 2014;12:12. doi: 10.1186/1479-5876-12-4. doi:10.1186/1479-5876-12-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kogure T, Lin W-L, Yan IK, Braconi C, Patel T. Inter-cellular nanovesicle-mediated microRNA transfer: a mechanism of environmental modulation of hepatocellular cancer cell growth. Hepatology. 2011;54(4):1237–1248. doi: 10.1002/hep.24504. doi:10.1002/hep.24504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kooijmans SA, Vader P, van Dommelen SM, van Solinge WW, Schiffelers RM. Exosome mimetics: a novel class of drug delivery systems. Int J Nanomed. 2012;7:1525–1541. doi: 10.2147/IJN.S29661. doi:10.2147/IJN.S29661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kooijmans SA, Stremersch S, Braeckmans K, de Smedt SC, Hendrix A, Wood MJ, Schiffelers RM, Raemdonck K, Vader P. Electroporation-induced siRNA precipitation obscures the efficiency of siRNA loading into extracellular vesicles. J Control Release. 2013;172(1):229–238. doi: 10.1016/j.jconrel.2013.08.014. doi:10.1016/j.jconrel.2013.08.014. [DOI] [PubMed] [Google Scholar]
- Koppers-Lalic D, Hogenboom MM, Middeldorp JM, Pegtel DM. Virus-modified exosomes for targeted RNA delivery; a new approach in nanomedicine. Adv Drug Deliv Rev. 2013;65(3):348–356. doi: 10.1016/j.addr.2012.07.006. doi:10.1016/j.addr.2012.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Korkut C, Ataman B, Ramachandran P, Ashley J, Barria R, Gherbesi N, Budnik V. Trans-synaptic transmission of vesicular Wnt signals through Evi/Wntless. Cell. 2009;139(2):393–404. doi: 10.1016/j.cell.2009.07.051. doi:10.1016/j.cell.2009.07.051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kosaka N, Iguchi H, Yoshioka Y, Takeshita F, Matsuki Y, Ochiya T. Secretory mechanisms and intercellular transfer of microRNAs in living cells. J Biol Chem. 2010;285(23):17442–17452. doi: 10.1074/jbc.M110.107821. doi:10.1074/jbc.M110.107821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kosaka N, Iguchi H, Yoshioka Y, Hagiwara K, Takeshita F, Ochiya T. Competitive interactions of cancer cells and normal cells via secretory microRNAs. J Biol Chem. 2012;287(2):1397–1405. doi: 10.1074/jbc.M111.288662. doi:10.1074/jbc.M111.288662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kosaka N, Takeshita F, Yoshioka Y, Hagiwara K, Katsuda T, Ono M, Ochiya T. Exosomal tumor-suppressive microRNAs as novel cancer therapy: “exocure” is another choice for cancer treatment. Adv Drug Deliv Rev. 2013;65(3):376–382. doi: 10.1016/j.addr.2012.07.011. doi:10.1016/j.addr.2012.07.011. [DOI] [PubMed] [Google Scholar]
- Kovacs GG, Budka H. Aging, the brain and human prion disease. Exp Gerontol. 2002;37(4):603–605. doi: 10.1016/s0531-5565(01)00219-4. doi:10.1016/S0531-5565(01)00219-4. [DOI] [PubMed] [Google Scholar]
- Kramer-Albers EM, Bretz N, Tenzer S, Winterstein C, Mobius W, Berger H, Nave KA, Schild H, Trotter J. Oligodendrocytes secrete exosomes containing major myelin and stress-protective proteins: Trophic support for axons? Proteomics Clin Appl. 2007;1(11):1446–1461. doi: 10.1002/prca.200700522. doi:10.1002/prca.200700522. [DOI] [PubMed] [Google Scholar]
- Krtolica A, Campisi J. Cancer and aging: a model for the cancer promoting effects of the aging stroma. Int J Biochem Cell Biol. 2002;34(11):1401–1414. doi: 10.1016/s1357-2725(02)00053-5. doi:10.1016/S1357-2725(02)00053-5. [DOI] [PubMed] [Google Scholar]
- Kruger S, AbdElmageed ZY, Hawke DH, Woerner PM, Jansen DA, Abdel-Mageed AB, Alt EU, Izadpanah R. Molecular characterization of exosome-like vesicles from breast cancer cells. BMC Cancer. 2014;14:14. doi: 10.1186/1471-2407-14-44. doi:10.1186/1471-2407-14-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kucharzewska P, Christianson HC, Welch JE, Svensson KJ, Fredlund E, Ringner M, Morgelin M, Bourseau-Guilmain E, Bengzon J, Belting M. Exosomes reflect the hypoxic status of glioma cells and mediate hypoxia-dependent activation of vascular cells during tumor development. Proc Natl Acad Sci USA. 2013;110(18):7312–7317. doi: 10.1073/pnas.1220998110. doi:10.1073/pnas.1220998110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuwabara Y, Ono K, Horie T, Nishi H, Nagao K, Kinoshita M, Watanabe S, Baba O, Kojima Y, Shizuta S, Imai M, Tamura T, Kita T, Kimura T. Increased microRNA-1 and microRNA-133a levels in serum of patients with cardiovascular disease indicate myocardial damage. Circ Cardiovasc Genet. 2011;4(4):446–454. doi: 10.1161/CIRCGENETICS.110.958975. doi:10.1161/circgenetics.110.958975. [DOI] [PubMed] [Google Scholar]
- Lachenal G, Pernet-Gallay K, Chivet M, Hemming FJ, Belly A, Bodon G, Blot B, Haase G, Goldberg Y, Sadoul R. Release of exosomes from differentiated neurons and its regulation by synaptic glutamatergic activity. Mol Cell Neurosci. 2011;46(2):409–418. doi: 10.1016/j.mcn.2010.11.004. doi:10.1016/j.mcn.2010.11.004. [DOI] [PubMed] [Google Scholar]
- Lai CP, Breakefield XO. Role of exosomes/microvesicles in the nervous system and use in emerging therapies. Front Physiol. 2012;3:228. doi: 10.3389/fphys.2012.00228. doi:10.3389/fphys.2012.00228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lai RC, Yeo RW, Tan KH, Lim SK. Exosomes for drug delivery—a novel application for the mesenchymal stem cell. Biotechnol Adv. 2013;31(5):543–551. doi: 10.1016/j.biotechadv.2012.08.008. doi:10.1016/j.biotechadv.2012.08.008. [DOI] [PubMed] [Google Scholar]
- Lakhal S, Wood MJ. Exosome nanotechnology: an emerging paradigm shift in drug delivery: exploitation of exosome nanovesicles for systemic in vivo delivery of RNAi heralds new horizons for drug delivery across biological barriers. BioEssays. 2011;33(10):737–741. doi: 10.1002/bies.201100076. doi:10.1002/bies.201100076. [DOI] [PubMed] [Google Scholar]
- Lancaster GI, Febbraio MA. Exosome-dependent trafficking of HSP70: a novel secretory pathway for cellular stress proteins. J Biol Chem. 2005;280(24):23349–23355. doi: 10.1074/jbc.M502017200. [DOI] [PubMed] [Google Scholar]
- Lao J, Madani J, Puertolas T, Alvarez M, Hernandez A, Pazo-Cid R, Artal A, Anton Torres A. Liposomal doxorubicin in the treatment of breast cancer patients: a review. J Drug Deliv. 2013;2013:456409. doi: 10.1155/2013/456409. doi:10.1155/2013/456409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lau C, Kim Y, Chia D, Spielmann N, Eibl G, Elashoff D, Wei F, Lin YL, Moro A, Grogan T, Chiang S, Feinstein E, Schafer C, Farrell J, Wong DT. Role of pancreatic cancer-derived exosomes in salivary biomarker development. J Biol Chem. 2013;288(37):26888–26897. doi: 10.1074/jbc.M113.452458. doi:10.1074/jbc.M113.452458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laulagnier K, Motta C, Hamdi S, Roy S, Fauvelle F, Pageaux JF, Kobayashi T, Salles JP, Perret B, Bonnerot C, Record M. Mast cell- and dendritic cell-derived exosomes display a specific lipid composition and an unusual membrane organization. Biochem J. 2004;380(Pt 1):161–171. doi: 10.1042/BJ20031594. doi:10.1042/BJ20031594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee JS, Feijen J. Polymersomes for drug delivery: design, formation and characterization. J Control Release. 2012;161(2):473–483. doi: 10.1016/j.jconrel.2011.10.005. doi:10.1016/j.jconrel.2011.10.005. [DOI] [PubMed] [Google Scholar]
- Lee RJ, Low PS. Delivery of liposomes into cultured Kb cells via folate receptor-mediated endocytosis. J Biol Chem. 1994;269(5):3198–3204. [PubMed] [Google Scholar]
- Lee TW, Matthews DA, Blair GE. Novel molecular approaches to cystic fibrosis gene therapy. Biochem J. 2005;387(Pt 1):1–15. doi: 10.1042/BJ20041923. doi:10.1042/BJ20041923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee HJ, Suk JE, Bae EJ, Lee JH, Paik SR, Lee SJ. Assembly-dependent endocytosis and clearance of extra-cellular alpha-synuclein. Int J Biochem Cell Biol. 2008;40(9):1835–1849. doi: 10.1016/j.biocel.2008.01.017. doi:10.1016/j.biocel.2008.01.017. [DOI] [PubMed] [Google Scholar]
- Lee HJ, Suk JE, Patrick C, Bae EJ, Cho JH, Rho S, Hwang D, Masliah E, Lee SJ. Direct transfer of alpha-synuclein from neuron to astroglia causes inflammatory responses in synucleinopathies. J Biol Chem. 2010;285(12):9262–9272. doi: 10.1074/jbc.M109.081125. doi:10.1074/jbc.M109.081125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee TH, D’Asti E, Magnus N, Al-Nedawi K, Meehan B, Rak J. Microvesicles as mediators of intercellular communication in cancer–the emerging science of cellular ‘debris’. Semin Immunopathol. 2011;33(5):455–467. doi: 10.1007/s00281-011-0250-3. doi:10.1007/s00281-011-0250-3. [DOI] [PubMed] [Google Scholar]
- Lee H, Lytton-Jean AK, Chen Y, Love KT, Park AI, Karagiannis ED, Sehgal A, Querbes W, Zurenko CS, Jayaraman M, Peng CG, Charisse K, Borodovsky A, Manoharan M, Donahoe JS, Truelove J, Nahrendorf M, Langer R, Anderson DG. Molecularly self-assembled nucleic acid nanoparticles for targeted in vivo siRNA delivery. Nat Nanotechnol. 2012a;7(6):389–393. doi: 10.1038/nnano.2012.73. doi:10.1038/nnano.2012.73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee JB, Hong J, Bonner DK, Poon Z, Hammond PT. Self-assembled RNA interference microsponges for efficient siRNA delivery. Nat Mater. 2012b;11(4):316–322. doi: 10.1038/nmat3253. doi:10.1038/nmat3253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee J-K, Park S-R, Jung B-K, Jeon Y-K, Lee Y-S, Kim M-K, Kim Y-G, Jang J-Y, Kim C-W. Exosomes derived from mesenchymal stem cells suppress angiogenesis by down-regulating VEGF expression in breast cancer cells. PLoS ONE. 2013;8(12):e84256. doi: 10.1371/journal.pone.0084256. doi:10.1371/journal.pone.0084256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lees AJ, Hardy J, Revesz T. Parkinson’s disease. Lancet. 2009;373(9680):2055–2066. doi: 10.1016/S0140-6736(09)60492-X. doi:10.1016/S0140-6736(09)60492-X. [DOI] [PubMed] [Google Scholar]
- Lespagnol A, Duflaut D, Beekman C, Blanc L, Fiucci G, Marine JC, Vidal M, Amson R, Telerman A. Exosome secretion, including the DNA damage-induced p53-dependent secretory pathway, is severely compromised in TSAP6/Steap3-null mice. Cell Death Differ. 2008;15(11):1723–1733. doi: 10.1038/cdd.2008.104. doi:10.1038/cdd.2008.104. [DOI] [PubMed] [Google Scholar]
- Levanen B, Bhakta NR, Paredes PT, Barbeau R, Hiltbrunner S, Pollack JL, Skold CM, Svartengren M, Grunewald J, Gabrielsson S, Eklund A, Larsson BM, Woodruff PG, Erle DJ, Wheelock AM. Altered microRNA profiles in bronchoalveolar lavage fluid exosomes in asthmatic patients. J Allergy Clin Immunol. 2013;131(3):894–903. doi: 10.1016/j.jaci.2012.11.039. doi:10.1016/j.jaci.2012.11.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leventis PA, Grinstein S. The distribution and function of phosphatidylserine in cellular membranes. Annu Rev Biophys. 2010;39:407–427. doi: 10.1146/annurev.biophys.093008.131234. doi:10.1146/annurev.biophys.093008.131234. [DOI] [PubMed] [Google Scholar]
- Levesque K, Halvorsen M, Abrahamyan L, Chatel-Chaix L, Poupon V, Gordon H, DesGroseillers L, Gatignol A, Mouland AJ. Trafficking of HIV-1 RNA is mediated by heterogeneous nuclear ribonucleoprotein A2 expression and impacts on viral assembly. Traffic. 2006;7(9):1177–1193. doi: 10.1111/j.1600-0854.2006.00461.x. doi:10.1111/j.1600-0854.2006.00461.x. [DOI] [PubMed] [Google Scholar]
- Levine DH, Ghoroghchian PP, Freudenberg J, Zhang G, Therien MJ, Greene MI, Hammer DA, Murali R. Polymer-somes: a new multi-functional tool for cancer diagnosis and therapy. Methods. 2008;46(1):25–32. doi: 10.1016/j.ymeth.2008.05.006. doi:10.1016/j.ymeth.2008.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J, Liu K, Liu Y, Xu Y, Zhang F, Yang H, Liu J, Pan T, Chen J, Wu M, Zhou X, Yuan Z. Exosomes mediate the cell-to-cell transmission of IFN-α-induced antiviral activity. Nat Immunol. 2013a;14:793–803. doi: 10.1038/ni.2647. doi:10.1038/ni.2647. [DOI] [PubMed] [Google Scholar]
- Li J, Zhang C, Li J, Fan L, Jiang X, Chen J, Pang Z, Zhang Q. Brain delivery of NAP with PEG-PLGA nanoparticles modified with phage display peptides. Pharm Res. 2013b;30(7):1813–1823. doi: 10.1007/s11095-013-1025-4. doi:10.1007/s11095-013-1025-4. [DOI] [PubMed] [Google Scholar]
- Liu C, Yu S, Zinn K, Wang J, Zhang L, Jia Y, Kappes JC, Barnes S, Kimberly RP, Grizzle WE, Zhang HG. Murine mammary carcinoma exosomes promote tumor growth by suppression of NK cell function. J Immunol. 2006;176(3):1375–1385. doi: 10.4049/jimmunol.176.3.1375. [DOI] [PubMed] [Google Scholar]
- Liu X-Q, Sun CY, Yang XZ, Wang J. Polymeric-micelle-based nanomedicine for siRNA delivery. Part Part Syst Charact. 2013;30(3):211–228. doi:10.1002/ppsc.201200061. [Google Scholar]
- Lodes MJ, Caraballo M, Suciu D, Munro S, Kumar A, Anderson B. Detection of cancer with serum miRNAs on an oligonucleotide microarray. PLoS One. 2009;4(7):e6229. doi: 10.1371/journal.pone.0006229. doi:10.1371/journal.pone.0006229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopez-Verrilli MA, Court FA. Transfer of vesicles from Schwann cells to axons: a novel mechanism of communication in the peripheral nervous system. Front Physiol. 2012;3:205. doi: 10.3389/fphys.2012.00205. doi:10.3389/fphys.2012.00205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopez-Verrilli MA, Picou F, Court FA. Schwann cell-derived exosomes enhance axonal regeneration in the peripheral nervous system. Glia. 2013;61(11):1795–1806. doi: 10.1002/glia.22558. doi:10.1002/glia.22558. [DOI] [PubMed] [Google Scholar]
- LoPresti C, Lomas H, Massignani M, Smart T, Battaglia G. Polymersomes: nature inspired nanometer sized compartments. J Mater Chem. 2009;19(22):3576–3590. doi:10.1039/B818869f. [Google Scholar]
- Lv LL, Cao YH, Ni HF, Xu M, Liu D, Liu H, Chen PS, Liu BC. MicroRNA-29c in urinary exosome/microvesicle as a biomarker of renal fibrosis. Am J Physiol-Renal Physiol. 2013;305(8):F1220–F1227. doi: 10.1152/ajprenal.00148.2013. doi:10.1152/ajprenal.00148.2013. [DOI] [PubMed] [Google Scholar]
- Mack M, Kleinschmidt A, Bruhl H, Klier C, Nelson PJ, Cihak J, Plachy J, Stangassinger M, Erfle V, Schlondorff D. Transfer of the chemokine receptor CCR5 between cells by membrane-derived microparticles: a mechanism for cellular human immunodeficiency virus 1 infection. Nat Med. 2000;6(7):769–775. doi: 10.1038/77498. doi:10.1038/77498. [DOI] [PubMed] [Google Scholar]
- Malmsten M. Inorganic nanomaterials as delivery systems for proteins, peptides, DNA, and siRNA. Curr Opin Colloid Int Sci. 2013;18(5):468–480. doi:10.1016/j.cocis.2013.06.002. [Google Scholar]
- Mangeot PE, Dollet S, Girard M, Ciancia C, Joly S, Peschanski M, Lotteau V. Protein transfer into human cells by VSV-G-induced nanovesicles. Mol Ther. 2011;19(9):1656–1666. doi: 10.1038/mt.2011.138. doi:10.1038/mt.2011.138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martinez MC, Freyssinet JM. Deciphering the plasma membrane hallmarks of apoptotic cells: phosphatidylserine transverse redistribution and calcium entry. BMC Cell Biol. 2001;2:20. doi: 10.1186/1471-2121-2-20. doi:10.1186/1471-2121-2-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martinez-Fong D, Bannon MJ, Trudeau LE, Gonzalez-Barrios JA, Arango-Rodriguez ML, Hernandez-Chan NG, Reyes-Corona D, Armendariz-Borunda J, Navarro-Quiroga I. NTS-Polyplex: a potential nanocarrier for neurotrophic therapy of Parkinson’s disease. Nanomedicine. 2012;8(7):1052–1069. doi: 10.1016/j.nano.2012.02.009. doi:10.1016/j.nano.2012.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marzesco AM, Janich P, Wilsch-Brauninger M, Dubreuil V, Langenfeld K, Corbeil D, Huttner WB. Release of extracellular membrane particles carrying the stem cell marker prominin-1 (CD133) from neural progenitors and other epithelial cells. J Cell Sci. 2005;118(Pt 13):2849–2858. doi: 10.1242/jcs.02439. doi:10.1242/jcs.02439. [DOI] [PubMed] [Google Scholar]
- Meckes DG, Jr, Shair KHY, Marquitz AR, Kung CP, Edwards RH, Raab-Traub N. Human tumor virus utilizes exosomes for intercellular communication. Proc Natl Acad Sci USA. 2010;107(47):20370–20375. doi: 10.1073/pnas.1014194107. doi:10.1073/pnas.1014194107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McConnell RE, Higginbotham JN, Shifrin DA, Jr, Tabb DL, Coffey RJ, Tyska MJ. The enterocyte microvillus is a vesicle-generating organelle. J Cell Biol. 2009;185(7):1285–1298. doi: 10.1083/jcb.200902147. doi:10.1083/jcb.200902147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Michael A, Bajracharya SD, Yuen PST, Zhou H, Star RA, Illei GG, Alevizos I. Exosomes from human saliva as a source of microRNA biomarkers. Oral Diseases. 2010;16(1):34–38. doi: 10.1111/j.1601-0825.2009.01604.x. doi:10.1111/j.1601-0825.2009.01604.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mills JC, Stone NL, Erhardt J, Pittman RN. Apoptotic membrane blebbing is regulated by myosin light chain phosphorylation. J Cell Biol. 1998;140(3):627–636. doi: 10.1083/jcb.140.3.627. doi:10.1083/jcb.140.3.627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minagar A, Jy W, Jimenez JJ, Sheremata WA, Mauro LM, Mao WW, Horstman LL, Ahn YS. Elevated plasma endothelial microparticles in multiple sclerosis. Neurology. 2001;56(10):1319–1324. doi: 10.1212/wnl.56.10.1319. [DOI] [PubMed] [Google Scholar]
- Mitchell PS, Parkin RK, Kroh EM, Fritz BR, Wyman SK, Pogosova-Agadjanyan EL, Peterson A, Noteboom J, O’Briant KC, Allen A, Lin DW, Urban N, Drescher CW, Knudsen BS, Stirewalt DL, Gentleman R, Vessella RL, Nelson PS, Martin DB, Tewari M. Circulating microRNAs as stable blood-based markers for cancer detection. Proc Natl Acad Sci USA. 2008;105(30):10513–10518. doi: 10.1073/pnas.0804549105. doi:10.1073/pnas.0804549105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moltzahn F, Olshen AB, Baehner L, Peek A, Fong L, Stoppler H, Simko J, Hilton JF, Carroll P, Blelloch R. Microfluidic-based multiplex qRT-PCR identifies diagnostic and prognostic microRNA signatures in the sera of prostate cancer patients. Cancer Res. 2011;71(2):550–560. doi: 10.1158/0008-5472.CAN-10-1229. doi:10.1158/0008-5472.can-10-1229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Montecalvo A, Larregina AT, Shufesky WJ, Beer Stolz D, Sullivan MLG, Karlsson JM, Baty CJ, Gibson GA, Erdos G, Wang Z, Milosevic J, Tkacheva OA, Divito SJ, Jordan R, Lyons-Weiler J, Watkins SC, Morelli AE. Mechanism of transfer of functional microRNAs between mouse dendritic cells via exosomes. Blood. 2012;119(3):756–766. doi: 10.1182/blood-2011-02-338004. doi:10.1182/blood-2011-02-338004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morel O, Morel N, Jesel L, Freyssinet JM, Toti F. Microparticles: a critical component in the nexus between inflammation, immunity, and thrombosis. Semin Immunopathol. 2011;33(5):469–486. doi: 10.1007/s00281-010-0239-3. doi:10.1007/s00281-010-0239-3. [DOI] [PubMed] [Google Scholar]
- Morelli AE, Larregina AT, Shufesky WJ. Endocytosis, intracellular sorting, and processing of exosomes by dendritic cells. Blood. 2004;104(10):3257–3266. doi: 10.1182/blood-2004-03-0824. doi:10.1182/blood-2004-03-0824. [DOI] [PubMed] [Google Scholar]
- Mourtas S, Canovi M, Zona C, Aurilia D, Niarakis A, La Ferla B, Salmona M, Nicotra F, Gobbi M, Antimisiaris SG. Curcumin-decorated nanoliposomes with very high affinity for amyloid-beta1-42 peptide. Biomaterials. 2011;32(6):1635–1645. doi: 10.1016/j.biomaterials.2010.10.027. doi:10.1016/j.biomaterials.2010.10.027. [DOI] [PubMed] [Google Scholar]
- Mueller G, Schneider M, Biemer-Daub G, Wied S. Microvesicles released from rat adipocytes and harboring glycosylphosphatidylinositol-anchored proteins transfer RNA stimulating lipid synthesis. Cell Signal. 2011;23(7):1207–1223. doi: 10.1016/j.cellsig.2011.03.013. doi:10.1016/j.cellsig.2011.03.013. [DOI] [PubMed] [Google Scholar]
- Mueller-Lantzsch N, Sauter M, Weiskircher A, Kramer K, Best B, Buck M, Grasser F. Human endogenous retroviral element K10 (HERV-K10) encodes a full-length gag homologous 73-kDa protein and a functional protease. AIDS Res Hum Retroviruses. 1993;9(4):343–350. doi: 10.1089/aid.1993.9.343. doi:10.1089/aid.1993.9.343. [DOI] [PubMed] [Google Scholar]
- Muhs A, Hickman DT, Pihlgren M, Chuard N, Giriens V, Meerschman C, van der Auwera I, van Leuven F, Sugawara M, Weingertner MC, Bechinger B, Greferath R, Kolonko N, Nagel-Steger L, Riesner D, Brady RO, Pfeifer A, Nicolau C. Liposomal vaccines with conformation-specific amyloid peptide antigens define immune response and efficacy in APP transgenic mice. Proc Natl Acad Sci USA. 2007;104(23):9810–9815. doi: 10.1073/pnas.0703137104. doi:10.1073/pnas.0703137104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munoz JL, Bliss SA, Greco SJ, Ramkissoon SH, Ligon KL, Rameshwar P. Delivery of functional anti-miR-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity. Mol Ther Nucleic Acids. 2013;2:e126. doi: 10.1038/mtna.2013.60. doi:10.1038/mtna.2013.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munro TP, Magee RJ, Kidd GJ, Carson JH, Barbarese E, Smith LM, Smith R. Mutational analysis of a heterogeneous nuclear ribonucleoprotein A2 response element for RNA trafficking. J Biol Chem. 1999;274(48):34389–34395. doi: 10.1074/jbc.274.48.34389. [DOI] [PubMed] [Google Scholar]
- Muralidharan-Chari V, Clancy J, Plou C, Romao M, Chavrier P, Raposo G, D’Souza-Schorey C. ARF6-regulated shedding of tumor cell-derived plasma membrane micro-vesicles. Curr Biol. 2009;19(22):1875–1885. doi: 10.1016/j.cub.2009.09.059. doi:10.1016/j.cub.2009.09.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mutlu NB, Degim Z, Yilmaz S, Essiz D, Nacar A. New perspective for the treatment of Alzheimer diseases: liposomal rivastigmine formulations. Drug Dev Ind Pharm. 2011;37(7):775–789. doi: 10.3109/03639045.2010.541262. doi:10.3109/03639045.2010.541262. [DOI] [PubMed] [Google Scholar]
- Nabhan JF, Hu R, Oh RS, Cohen SN, Lu Q. Formation and release of arrestin domain-containing protein 1-mediated microvesicles (ARMMs) at plasma membrane by recruitment of TSG101 protein. Proc Natl Acad Sci USA. 2012;109(11):4146–4151. doi: 10.1073/pnas.1200448109. doi:10.1073/pnas.1200448109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen DG, Booth A, Gould SJ, Hildreth JEK. Evidence that HIV budding in primary macrophages occurs through the exosome release pathway. J Biol Chem. 2003;278(52):52347–52354. doi: 10.1074/jbc.M309009200. doi:10.1074/jbc.M309009200. [DOI] [PubMed] [Google Scholar]
- Nicolau C, Greferath R, Balaban T, Lazarte J, Hopkins R. A liposome-based therapeutic vaccine against beta-amyloid plaques on the pancreas of transgenic NORBA mice. Proc Natl Acad Sci USA. 2002;99(4):2332–2337. doi: 10.1073/pnas.022627199. doi:10.1073/pnas.022627199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Noble GT, Stefanick JF, Ashley JD, Kiziltepe T, Bilgicer B. Ligand-targeted liposome design: challenges and fundamental considerations. Trends Biotechnol. 2014;32(1):32–45. doi: 10.1016/j.tibtech.2013.09.007. doi:10.1016/j.tibtech.2013.09.007. [DOI] [PubMed] [Google Scholar]
- Nolte-’t Hoen ENM, Buschow SI, Anderton SM, Stoorvogel W, Wauben MHM. Activated T cells recruit exosomes secreted by dendritic cells via LFA-1. Blood. 2009;113(9):1977–1981. doi: 10.1182/blood-2008-08-174094. doi:10.1182/blood-2008-08-174094. [DOI] [PubMed] [Google Scholar]
- Nolte-’t Hoen EN, van der Vlist EJ, Aalberts M, Mertens HC, Bosch BJ, Bartelink W, Mastrobattista E, van Gaal EV, Stoorvogel W, Arkesteijn GJ, Wauben MH. Quantitative and qualitative flow cytometric analysis of nano-sized cell-derived membrane vesicles. Nanomedicine. 2012;8(5):712–720. doi: 10.1016/j.nano.2011.09.006. doi:10.1016/j.nano.2011.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ogata-Kawata H, Izumiya M, Kurioka D, Honma Y, Yamada Y, Furuta K, Gunji T, Ohta H, Okamoto H, Sonoda H, Watanabe M, Nakagama H, Yokota J, Kohno T, Tsuchiya N. Circulating exosomal microRNAs as biomarkers of colon cancer. PLoSOne. 2014;9(4):9. doi: 10.1371/journal.pone.0092921. doi:10.1371/journal.pone.0092921. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohno S, Takanashi M, Sudo K, Ueda S, Ishikawa A, Matsuyama N, Fujita K, Mizutani T, Ohgi T, Ochiya T, Gotoh N, Kuroda M. Systemically injected exosomes targeted to EGFR deliver antitumor microRNA to breast cancer cells. Mol Ther. 2013;21(1):185–191. doi: 10.1038/mt.2012.180. doi:10.1038/mt.2012.180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohshima K, Inoue K, Fujiwara A, Hatakeyama K, Kanto K, Watanabe Y, Muramatsu K, Fukuda Y, Ogura S, Yamaguchi K, Mochizuki T. Let-7 microRNA family is selectively secreted into the extracellular environment via exosomes in a metastatic gastric cancer cell line. PLoS One. 2010;5(10):e13247. doi: 10.1371/journal.pone.0013247. doi:10.1371/journal.pone.0013247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Olson SD, Kambal A, Pollock K, Mitchell GM, Stewart H, Kalomoiris S, Cary W, Nacey C, Pepper K, Nolta JA. Examination of mesenchymal stem cell-mediated RNAi transfer to Huntington’s disease affected neuronal cells for reduction of huntingtin. Mol Cell Neurosci. 2012;49(3):271–281. doi: 10.1016/j.mcn.2011.12.001. doi:10.1016/j.mcn.2011.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ostrowski M, Carmo NB, Krumeich S, Fanget I, Raposo G, Savina A, Moita CF, Schauer K, Hume AN, Freitas RP, Goud B, Benaroch P, Hacohen N, Fukuda M, Desnos C, Seabra MC, Darchen F, Amigorena S, Moita LF, Thery C. Rab27a and Rab27b control different steps of the exosome secretion pathway. Nat Cell Biol. 2010;12(1):19–30. doi: 10.1038/ncb2000. doi:10.1038/ncb2000. sup pp 11-13. [DOI] [PubMed] [Google Scholar]
- Palma J, Yaddanapudi SC, Pigati L, Havens MA, Jeong S, Weiner GA, Weimer KME, Stern B, Hastings ML, Duelli DM. MicroRNAs are exported from malignant cells in customized particles. Nucleic Acids Res. 2012;40(18):9125–9138. doi: 10.1093/nar/gks656. doi:10.1093/nar/gks656. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pang Z, Lu W, Gao H, Hu K, Chen J, Zhang C, Gao X, Jiang X, Zhu C. Preparation and brain delivery property of biodegradable polymersomes conjugated with OX26. J Control Release. 2008;128(2):120–127. doi: 10.1016/j.jconrel.2008.03.007. doi:10.1016/j.jconrel.2008.03.007. [DOI] [PubMed] [Google Scholar]
- Pang Z, Feng L, Hua R, Chen J, Gao H, Pan S, Jiang X, Zhang P. Lactoferrin-conjugated biodegradable polymersome holding doxorubicin and tetrandrine for chemotherapy of glioma rats. Mol Pharm. 2010;7(6):1995–2005. doi: 10.1021/mp100277h. doi:10.1021/mp100277h. [DOI] [PubMed] [Google Scholar]
- Pang Z, Gao H, Yu Y, Guo L, Chen J, Pan S, Ren J, Wen Z, Jiang X. Enhanced intracellular delivery and chemotherapy for glioma rats by transferrin-conjugated biodegradable polymersomes loaded with doxorubicin. Bioconjug Chem. 2011;22(6):1171–1180. doi: 10.1021/bc200062q. doi:10.1021/bc200062q. [DOI] [PubMed] [Google Scholar]
- Pangburn T, Georgiou K, Bates F, Kokkoli E. Targeted polymersome delivery of siRNA induces cell death of breast cancer cells dependent upon Orai3 protein expression. Langmuir. 2012;28(35):12816–12830. doi: 10.1021/la300874z. doi:10.1021/la300874z. [DOI] [PubMed] [Google Scholar]
- Park J, Tan H, Datta A, Lai R, Zhang H, Meng W, Lim S, Sze S. Hypoxic tumor cell modulates its microenvironment to enhance angiogenic and metastatic potential by secretion of proteins and exosomes. Mol Cell Prot. 2010;9(6):1085–1099. doi: 10.1074/mcp.M900381-MCP200. doi:10.1074/mcp.M900381-MCP200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Patel DM, Arnold PY, White GA, Nardella JP, Mannie MD. Class II MHC/peptide complexes are released from APC and are acquired by T cell responders during specific antigen recognition. J Immunol. 1999;163(10):5201–5210. [PubMed] [Google Scholar]
- Patel DA, Henry JE, Good TA. Attenuation of beta-amyloid-induced toxicity by sialic-acid-conjugated dendrimers: role of sialic acid attachment. Brain Res. 2007;1161:95–105. doi: 10.1016/j.brainres.2007.05.055. doi:10.1016/j.brainres.2007.05.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paula-Barbosa M, Mota Cardoso R, Faria R, Cruz C. Multivesicular bodies in cortical dendrites of two patients with Alzheimer’s disease. J Neurol Sci. 1978;36(2):259–264. doi: 10.1016/0022-510x(78)90086-2. doi:10.1016/0022-510X(78)90086-2. [DOI] [PubMed] [Google Scholar]
- Pegtel DM, Cosmopoulos K, Thorley-Lawson DA, van Eijndhoven MAJ, Hopmans ES, Lindenberg JL, de Gruijl TD, Wurdinger T, Middeldorp JM. Functional delivery of viral miRNAs via exosomes. Proc Natl Acad Sci USA. 2010;107(14):6328–6333. doi: 10.1073/pnas.0914843107. doi:10.1073/pnas.0914843107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pelloski CE, Ballman KV, Furth AF, Zhang L, Lin E, Sulman EP, Bhat K, McDonald JM, Yung WK, Colman H, Woo SY, Heimberger AB, Suki D, Prados MD, Chang SM, Barker FG, II, Buckner JC, James CD, Aldape K. Epidermal growth factor receptor variant III status defines clinically distinct subtypes of glioblastoma. J Clin Oncol. 2007;25(16):2288–2294. doi: 10.1200/JCO.2006.08.0705. doi:10.1200/jco.2006.08.0705. [DOI] [PubMed] [Google Scholar]
- Piccoli G, Condliffe SB, Bauer M, Giesert F, Boldt K, De Astis S, Meixner A, Sarioglu H, Vogt-Weisenhorn DM, Wurst W, Gloeckner CJ, Matteoli M, Sala C, Ueffing M. LRRK2 controls synaptic vesicle storage and mobilization within the recycling pool. J Neurosci. 2011;31(6):2225–2237. doi: 10.1523/JNEUROSCI.3730-10.2011. doi:10.1523/JNEUROSCI.3730-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pincetic A, Leis J. The mechanism of budding of retro-viruses from cell membranes. Adv Virol. 2009;2009:623969. doi: 10.1155/2009/623969. doi:10.1155/2009/623969. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piper RC, Katzmann DJ. Biogenesis and function of multivesicular bodies. Annu Rev Cell Dev Biol. 2007;23:519–547. doi: 10.1146/annurev.cellbio.23.090506.123319. doi:10.1146/annurev.cellbio.23.090506.123319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Potolicchio I, Carven GJ, Xu X, Stipp C, Riese RJ, Stern LJ, Santambrogio L. Proteomic analysis of microglia-derived exosomes: metabolic role of the aminopeptidase CD13 in neuropeptide catabolism. J Immunol. 2005;175(4):2237–2243. doi: 10.4049/jimmunol.175.4.2237. [DOI] [PubMed] [Google Scholar]
- Pourtau L, Oliveira H, Thevenot J, Wan Y, Brisson A, Sandre O, Miraux S, Thiaudiere E, Lecommandoux S. Antibody-functionalized magnetic polymersomes: in vivo targeting and imaging of bone metastases using high resolution MRI. Adv Healthcare Mater. 2013;2(11):1420–1424. doi: 10.1002/adhm.201300061. doi:10.1002/adhm.201300061. [DOI] [PubMed] [Google Scholar]
- Principe S, Hui A, Bruce J, Sinha A, Liu F-F, Kislinger T. Tumor-derived exosomes and microvesicles in head and neck cancer: implications for tumor biology and biomarker discovery. Proteomics. 2013;13(10-11):1608–1623. doi: 10.1002/pmic.201200533. doi:10.1002/pmic.201200533. [DOI] [PubMed] [Google Scholar]
- Pusic AD, Kraig RP. Youth and environmental enrichment generate serum exosomes containing miR-219 that promote CNS myelination. Glia. 2014;62(2):284–299. doi: 10.1002/glia.22606. doi:10.1002/glia.22606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pusic AD, Pusic KM, Clayton BL, Kraig RP. IFNγ-stimulated dendritic cell exosomes as a potential therapeutic for remyelination. J Neuroimmunol. 2014;266(1-2):12–23. doi: 10.1016/j.jneuroim.2013.10.014. doi:10.1016/j.jneuroim.2013.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Putz U, Howitt J, Lackovic J, Foot N, Kumar S, Silke J, Tan SS. Nedd4 family-interacting protein 1 (Ndfip1) is required for the exosomal secretion of Nedd4 family proteins. J Biol Chem. 2008;283(47):32621–32627. doi: 10.1074/jbc.M804120200. doi:10.1074/jbc.M804120200. [DOI] [PubMed] [Google Scholar]
- Qu Y, Franchi L, Nunez G, Dubyak GR. Nonclassical IL-1 beta secretion stimulated by P2X7 receptors is dependent on inflammasome activation and correlated with exosome release in murine macrophages. J Immunol. 2007;179(3):1913–1925. doi: 10.4049/jimmunol.179.3.1913. [DOI] [PubMed] [Google Scholar]
- Querfurth H, LaFerla F. Alzheimer’s disease. N Engl J Med. 2010;362:329–344. doi: 10.1056/NEJMra0909142. [DOI] [PubMed] [Google Scholar]
- Rai K, Takigawa N, Ito S, Kashihara H, Ichihara E, Yasuda T, Shimizu K, Tanimoto M, Kiura K. Liposomal delivery of microRNA-7-expressing plasmid overcomes epidermal growth factor receptor tyrosine kinase inhibitor-resistance in lung cancer cells. Mol Cancer Ther. 2011;10(9):1720–1727. doi: 10.1158/1535-7163.MCT-11-0220. doi:10.1158/1535-7163.MCT-11-0220. [DOI] [PubMed] [Google Scholar]
- Rajendra KS, Hae-Won K. Inorganic nanobiomaterial drug carriers for medicine. Tissue Eng Regenerative Med. 2013;10(6):296–309. doi:10.1007/s13770-013-1092-y. [Google Scholar]
- Rajendran L, Honsho M, Zahn T. Alzheimer’s disease β-amyloid peptides are released in association with exosomes. Proc Natl Acad Sci USA. 2006;103(30):11172–11177. doi: 10.1073/pnas.0603838103. doi:10.1073/pnas.0603838103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rana S, Malinowska K, Zöller M. Exosomal tumor microRNA modulates premetastatic organ cells. Neoplasia. 2013;15(3):281–295. doi: 10.1593/neo.122010. doi:10.1593/neo.122010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raposo G, Stoorvogel W. Extracellular vesicles: exosomes, microvesicles, and friends. J Cell Biol. 2013;200(4):373–383. doi: 10.1083/jcb.201211138. doi:10.1083/jcb.201211138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raposo G, Nijman HW, Stoorvogel W, Liejendekker R, Harding CV, Melief CJ, Geuze HJ. B lymphocytes secrete antigen-presenting vesicles. J Exp Med. 1996;183(3):1161–1172. doi: 10.1084/jem.183.3.1161. doi:10.1084/jem.183.3.1161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ratajczak J, Miekus K, Kucia M, Zhang J, Reca R, Dvorak P, Ratajczak MZ. Embryonic stem cell-derived microvesicles reprogram hematopoietic progenitors: evidence for horizontal transfer of mRNA and protein delivery. Leukemia. 2006;20(5):847–856. doi: 10.1038/sj.leu.2404132. doi:10.1038/sj.leu.2404132. [DOI] [PubMed] [Google Scholar]
- Reiche J, Pauli G, Ellerbrok H. Differential expression of human endogenous retrovirus K transcripts in primary human melanocytes and melanoma cell lines after UV irradiation. Melanoma Res. 2010;20(5):435–440. doi: 10.1097/CMR.0b013e32833c1b5d. doi:10.1097/CMR.0b013e32833c1b5d. [DOI] [PubMed] [Google Scholar]
- Rijcken C, Soga O, Hennink W, van Nostrum C. Triggered destabilisation of polymeric micelles and vesicles by changing polymers polarity: an attractive tool for drug delivery. J Control Release. 2007;120(3):131–148. doi: 10.1016/j.jconrel.2007.03.023. doi:10.1016/j.jconrel.2007.03.023. [DOI] [PubMed] [Google Scholar]
- Roberts RL, Fine RE, Sandra A. Receptor-mediated endocytosis of transferrin at the blood-brain barrier. J Cell Sci. 1993;104(Pt 2):521–532. doi: 10.1242/jcs.104.2.521. [DOI] [PubMed] [Google Scholar]
- Robertson C, Booth SA, Beniac DR, Coulthart MB, Booth TF, McNicol A. Cellular prion protein is released on exosomes from activated platelets. Blood. 2006;107(10):3907–3911. doi: 10.1182/blood-2005-02-0802. doi:10.1182/blood-2005-02-0802. [DOI] [PubMed] [Google Scholar]
- Roh Y, Lee J, Kiatwuthinon P, Hartman M, Cha J, Um S, Muller D, Luo D. DNAsomes: multifunctional DNA-based nanocarriers. Small. 2011;7(1):74–78. doi: 10.1002/smll.201000752. doi:10.1002/smll.201000752. [DOI] [PubMed] [Google Scholar]
- Russo I, Bubacco L, Greggio E. Exosomes-associated neurodegeneration and progression of Parkinson’s disease. Am J Neurodegen Dis. 2012;1(3):217–225. [PMC free article] [PubMed] [Google Scholar]
- Saksena S, Sun J, Chu T, Emr S. ESCRTing proteins in the endocytic pathway. Trends Biochem Sci. 2007;32(12):561–573. doi: 10.1016/j.tibs.2007.09.010. doi:10.1016/j.tibs.2007.09.010. [DOI] [PubMed] [Google Scholar]
- Saman S, Kim W, Raya M, Visnick Y, Miro S, Saman S, Jackson B, McKee AC, Alvarez VE, Lee NC, Hall GF. Exosome-associated tau is secreted in tauopathy models and is selectively phosphorylated in cerebrospinal fluid in early Alzheimer disease. J Biol Chem. 2012;287(6):3842–3849. doi: 10.1074/jbc.M111.277061. doi:10.1074/jbc.M111.277061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sawant RR, Jhaveri AM, Torchilin VP. Immunomicelles for advancing personalized therapy. Adv Drug Deliv Rev. 2012;64(13):1436–1446. doi: 10.1016/j.addr.2012.08.003. doi:10.1016/j.addr.2012.08.003. [DOI] [PubMed] [Google Scholar]
- Schiera G, Proia P, Alberti C, Mineo M, Savettieri G, Di Liegro I. Neurons produce FGF2 and VEGF and secrete them at least in part by shedding extracellular vesicles. J Cell Mol Med. 2007;11(6):1384–1394. doi: 10.1111/j.1582-4934.2007.00100.x. doi:10.1111/j.1582-4934.2007.00100.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schneider A, Simons M. Exosomes: vesicular carriers for intercellular communication in neurodegenerative disorders. Cell Tissue Res. 2013;352(1):33–47. doi: 10.1007/s00441-012-1428-2. doi:10.1007/s00441-012-1428-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sebbagh M, Renvoize C, Hamelin J, Riche N, Bertoglio J, Breard J. Caspase-3-mediated cleavage of ROCK I induces MLC phosphorylation and apoptotic membrane blebbing. Nat Cell Biol. 2001;3(4):346–352. doi: 10.1038/35070019. doi:10.1038/35070019. [DOI] [PubMed] [Google Scholar]
- Sekhon BS, Kamboj SR. Inorganic nanomedicine—Part 1. Nanomedicine. 2010a;6(4):516–522. doi: 10.1016/j.nano.2010.04.004. doi:10.1016/j.nano.2010.04.004. [DOI] [PubMed] [Google Scholar]
- Sekhon BS, Kamboj SR. Inorganic nanomedicine—Part 2. Nanomedicine. 2010b;6(5):612–618. doi: 10.1016/j.nano.2010.04.003. doi:10.1016/j.nano.2010.04.003. [DOI] [PubMed] [Google Scholar]
- Sharma P, Schiapparelli L, Cline HT. Exosomes function in cell-cell communication during brain circuit development. Curr Opin Neurobiol. 2013;23(6):997–1004. doi: 10.1016/j.conb.2013.08.005. doi:10.1016/j.conb.2013.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin N, Jeong H, Kwon J, Heo HY, Kwon JJ, Yun HJ, Kim CH, Han BS, Tong Y, Shen J, Hatano T, Hattori N, Kim KS, Chang S, Seol W. LRRK2 regulates synaptic vesicle endocytosis. Exp Cell Res. 2008;314(10):2055–2065. doi: 10.1016/j.yexcr.2008.02.015. doi:10.1016/j.yexcr.2008.02.015. [DOI] [PubMed] [Google Scholar]
- Shin-ichiro O, Akio I, Masahiko K. Roles of exosomes and microvesicles in disease pathogenesis. Adv Drug Del Rev. 2013;65(3):398–401. doi: 10.1016/j.addr.2012.07.019. doi:10.1016/j.addr.2012.07.019. [DOI] [PubMed] [Google Scholar]
- Shu D, Shu Y, Haque F, Abdelmawla S, Guo P. Thermodynamically stable RNA three-way junction for constructing multifunctional nanoparticles for delivery of therapeutics. Nat Nanotechnol. 2011;6(10):658–667. doi: 10.1038/nnano.2011.105. doi:10.1038/nnano.2011.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shu Y, Haque F, Shu D, Li W, Zhu Z, Kotb M, Lyubchenko Y, Guo P. Fabrication of 14 different RNA nanoparticles for specific tumor targeting without accumulation in normal organs. RNA. 2013a;19(6):767–777. doi: 10.1261/rna.037002.112. doi:10.1261/rna.037002.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shu Y, Shu D, Haque F, Guo P. Fabrication of pRNA nanoparticles to deliver therapeutic RNAs and bioactive compounds into tumor cells. Nat Protoc. 2013b;8(9):1635–1659. doi: 10.1038/nprot.2013.097. doi:10.1038/nprot.2013.097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shu Y, Pi F, Sharma A, Rajabi M, Haque F, Shu D, Leggas M, Evers BM, Guo P. Stable RNA nanoparticles as potential new generation drugs for cancer therapy. Adv Drug Del Rev. 2014;66:74–89. doi: 10.1016/j.addr.2013.11.006. doi:10.1016/j.addr.2013.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silva J, Garcia V, Zaballos A, Provencio M, Lombardia L, Almonacid L, Garcia JM, Dominguez G, Pena C, Diaz R, Herrera M, Varela A, Bonilla F. Vesicle-related microRNAs in plasma of nonsmall cell lung cancer patients and correlation with survival. Eur Respir J. 2011;37(3):617–623. doi: 10.1183/09031936.00029610. doi:10.1183/09031936.00029610. [DOI] [PubMed] [Google Scholar]
- Simak J, Gelderman MP, Yu H, Wright V, Baird AE. Circulating endothelial microparticles in acute ischemic stroke: a link to severity, lesion volume and outcome. J Thromb Haemost. 2006;4(6):1296–1302. doi: 10.1111/j.1538-7836.2006.01911.x. doi:10.1111/j.1538-7836.2006.01911.x. [DOI] [PubMed] [Google Scholar]
- Simona F, Laura S, Simona T, Riccardo A. Contribution of proteomics to understanding the role of tumor-derived exosomes in cancer progression: state of the art and new perspectives. Proteomics. 2013;13(10-11):1581–1594. doi: 10.1002/pmic.201200398. doi:10.1002/pmic.201200398. [DOI] [PubMed] [Google Scholar]
- Simons M, Raposo G. Exosomes—vesicular carriers for intercellular communication. Curr Opin Cell Biol. 2009;21(4):575–581. doi: 10.1016/j.ceb.2009.03.007. doi:10.1016/j.ceb.2009.03.007. [DOI] [PubMed] [Google Scholar]
- Skog J, Würdinger T, van Rijn S, Meijer D, Gainche L, Sena-Esteves M, Curry W, Carter B, Krichevsky A, Breakefield X. Glioblastoma microvesicles transport RNA and proteins that promote tumour growth and provide diagnostic biomarkers. Nat Cell Biol. 2008;10(12):1470–1476. doi: 10.1038/ncb1800. doi:10.1038/ncb1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soekmadji C, Russell P, Nelson C. Exosomes in prostate cancer: putting together the pieces of a puzzle. Cancers. 2013;5(4):1522–1544. doi: 10.3390/cancers5041522. doi:10.3390/cancers5041522. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Son SJ, Bai X, Lee SB. Inorganic hollow nanoparticles and nanotubes in nanomedicine Part 1. Drug/gene delivery applications. Drug Discov Today. 2007a;12(15-16):650–656. doi: 10.1016/j.drudis.2007.06.002. doi:10.1016/j.drudis.2007.06.002. [DOI] [PubMed] [Google Scholar]
- Son SJ, Bai X, Lee SB. Inorganic hollow nanoparticles and nanotubes in nanomedicine Part 2: Imaging, diagnostic, and therapeutic applications. Drug Discov Today. 2007b;12(15-16):657–663. doi: 10.1016/j.drudis.2007.06.012. doi:10.1016/j.drudis.2007.06.012. [DOI] [PubMed] [Google Scholar]
- Soo CY, Song Y, Zheng Y, Campbell EC, Riches AC, Gunn-Moore F, Powis SJ. Nanoparticle tracking analysis monitors microvesicle and exosome secretion from immune cells. Immunology. 2012;136(2):192–197. doi: 10.1111/j.1365-2567.2012.03569.x. doi:10.1111/j.1365-2567.2012.03569.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spagnou S, Miller A, Keller M. Lipidic carriers of siR-NA: differences in the formulation, cellular uptake, and delivery with plasmid DNA. Biochemistry. 2004;43(42):13348–13356. doi: 10.1021/bi048950a. doi:10.1021/bi048950a. [DOI] [PubMed] [Google Scholar]
- Spuch C, Navarro C. Liposomes for targeted delivery of active agents against neurodegenerative diseases (Alzheimer’s disease and Parkinson’s disease) J Drug Deliv. 2011;2011:469679. doi: 10.1155/2011/469679. doi:10.1155/2011/469679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Staruch R, Chopra R, Hynynen K. Localised drug release using MRI-controlled focused ultrasound hyperthermia. Int J Hypertherm. 2011;27(2):156–171. doi: 10.3109/02656736.2010.518198. doi:10.3109/02656736.2010.518198. [DOI] [PubMed] [Google Scholar]
- Stoll E, Horner P, Rostomily R. The impact of age on oncogenic potential: tumor-initiating cells and the brain microenvironment. Aging Cell. 2013;12(5):733–741. doi: 10.1111/acel.12104. doi:10.1111/acel.12104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Subra C, Laulagnier K, Perret B, Record M. Exosome lipidomics unravels lipid sorting at the level of multivesicular bodies. Biochimie. 2007;89(2):205–212. doi: 10.1016/j.biochi.2006.10.014. doi:10.1016/j.biochi.2006.10.014. [DOI] [PubMed] [Google Scholar]
- Sun D, Zhuang X, Xiang X, Liu Y, Zhang S, Liu C, Barnes S, Grizzle W, Miller D, Zhang H-G. A novel nano-particle drug delivery system: the anti-inflammatory activity of curcumin is enhanced when encapsulated in exosomes. Mol Ther. 2010;18(9):1606–1614. doi: 10.1038/mt.2010.105. doi:10.1038/mt.2010.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Svensson K, Kucharzewska P, Christianson H, Sköld S, Löfstedt T, Johansson M, Mörgelin M, Bengzon J, Ruf W, Belting M. Hypoxia triggers a proangiogenic pathway involving cancer cell microvesicles and PAR-2-mediated heparin-binding EGF signaling in endothelial cells. Proc Natl Acad Sci USA. 2011;108(32):13147–13152. doi: 10.1073/pnas.1104261108. doi:10.1073/pnas.1104261108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szajnik M, Czystowska M, Szczepanski MJ, Mandapathil M, Whiteside TL. Tumor-derived microvesicles induce, expand and up-regulate biological activities of human regulatory T cells (Treg) PLoS One. 2010;5(7):e11469. doi: 10.1371/journal.pone.0011469. doi:10.1371/journal.pone.0011469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tadokoro H, Umezu T, Ohyashiki K, Hirano T, Ohyashiki JH. Exosomes derived from hypoxic leukemia cells enhance tube formation in endothelial cells. J Biol Chem. 2013;288(48):34343–34351. doi: 10.1074/jbc.M113.480822. doi:10.1074/jbc.M113.480822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tai-Kin W, Nicolau C, Hofschneider PH. Appearance of β-lactamase activity in animal cells upon liposome-mediated gene transfer. Gene. 1980;10(2):87–94. doi: 10.1016/0378-1119(80)90126-2. doi:10.1016/0378-1119(80)90126-2. [DOI] [PubMed] [Google Scholar]
- Takahashi RH, Milner TA, Li F, Nam EE, Edgar MA, Yamaguchi H, Beal MF, Xu H, Greengard P, Gouras GK. Intraneuronal Alzheimer Aβ42 accumulates in multivesicular bodies and is associated with synaptic pathology. Am J Pathol. 2002;161(5):1869–1879. doi: 10.1016/s0002-9440(10)64463-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tamboli IY, Barth E, Christian L, Siepmann M, Kumar S, Singh S, Tolksdorf K, Heneka MT, Lutjohann D, Wunderlich P, Walter J. Statins promote the degradation of extra-cellular amyloid {beta}-peptide by microglia via stimulation of exosome-associated insulin-degrading enzyme (IDE) secretion. J Biol Chem. 2010;285(48):37405–37414. doi: 10.1074/jbc.M110.149468. doi:10.1074/jbc.M110.149468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan A, De La Pena H, Seifalian AM. The application of exosomes as a nanoscale cancer vaccine. Int J Nanomed. 2010;5:889–900. doi: 10.2147/IJN.S13402. doi:10.2147/IJN.S13402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan A, Rajadas J, Seifalian A. Exosomes as nano-theranostic delivery platforms for gene therapy. Adv Drug Del Rev. 2013;65(3):357–367. doi: 10.1016/j.addr.2012.06.014. doi:10.1016/j.addr.2012.06.014. [DOI] [PubMed] [Google Scholar]
- Tang S, Martinez LJ, Sharma A, Chai M. Synthesis and characterization of water-soluble and photostable l-DOPA dendrimers. Org Lett. 2006;8(20):4421–4424. doi: 10.1021/ol061449l. doi:10.1021/ol061449l. [DOI] [PubMed] [Google Scholar]
- Taruscio D, Mantovani A. Factors regulating endogenous retroviral sequences in human and mouse. Cytogenet Genome Res. 2004;105(2-4):351–362. doi: 10.1159/000078208. doi:10.1159/000078208. [DOI] [PubMed] [Google Scholar]
- Taylor DD, Gercel-Taylor C. MicroRNA signatures of tumor-derived exosomes as diagnostic biomarkers of ovarian cancer. Gynecol Oncol. 2008;110(1):13–21. doi: 10.1016/j.ygyno.2008.04.033. doi:10.1016/j.ygyno.2008.04.033. [DOI] [PubMed] [Google Scholar]
- Taylor DD, Akyol S, Gercel-Taylor C. Pregnancy-associated exosomes and their modulation of T cell signaling. J Immunol. 2006;176(3):1534–1542. doi: 10.4049/jimmunol.176.3.1534. [DOI] [PubMed] [Google Scholar]
- Taylor AR, Robinson MB, Gifondorwa DJ, Tytell M, Milligan CE. Regulation of heat shock protein 70 release in astrocytes: role of signaling kinases. Dev Neurobiol. 2007;67(13):1815–1829. doi: 10.1002/dneu.20559. doi:10.1002/dneu.20559. [DOI] [PubMed] [Google Scholar]
- Taylor M, Moore S, Mourtas S, Niarakis A, Re F, Zona C, La Ferla B, Nicotra F, Masserini M, Antimisiaris SG, Gregori M, Allsop D. Effect of curcumin-associated and lipid ligand-functionalized nanoliposomes on aggregation of the Alzheimer’s Abeta peptide. Nanomedicine. 2011;7(5):541–550. doi: 10.1016/j.nano.2011.06.015. doi:10.1016/j.nano.2011.06.015. [DOI] [PubMed] [Google Scholar]
- Templeton NS. Cationic liposome-mediated gene delivery in vivo. Biosci Rep. 2002;22(2):283–295. doi: 10.1023/a:1020142823595. doi:10.1023/A:1020142823595. [DOI] [PubMed] [Google Scholar]
- Tennyson LD, Clemens B. The unique role of nanoparticles in nanomedicine: imaging, drug delivery and therapy. Chem Soc Rev. 2012;41(7):2885–2911. doi: 10.1039/c2cs15260f. doi:10.1039/c2cs15260f. [DOI] [PubMed] [Google Scholar]
- Theresa MA, Pieter RC. Liposomal drug delivery systems: from concept to clinical applications. Adv Drug Del Rev. 2013;65(1):36–48. doi: 10.1016/j.addr.2012.09.037. doi:10.1016/j.addr.2012.09.037. [DOI] [PubMed] [Google Scholar]
- Thery C. Exosomes: secreted vesicles and intercellular communications. Biol Rep. 2011;3:1–5. doi: 10.3410/B3-15. doi:10.3410/b3-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Théry C, Duban L, Segura E, Véron P, Lantz O, Amigorena S. Indirect activation of naïve CD4 + T cells by dendritic cell-derived exosomes. Nat Immunol. 2002a;3(12):1156–1162. doi: 10.1038/ni854. [DOI] [PubMed] [Google Scholar]
- Théry C, Zitvogel L, Amigorena S. Exosomes: composition, biogenesis and function. Nat Rev Immunol. 2002b;2(8):569–579. doi: 10.1038/nri855. [DOI] [PubMed] [Google Scholar]
- Théry C, Ostrowski M, Segura E. Membrane vesicles as conveyors of immune responses. Nat Rev Immunol. 2009;9(8):581–593. doi: 10.1038/nri2567. doi:10.1038/nri2567. [DOI] [PubMed] [Google Scholar]
- Theunis C, Crespo-Biel N, Gafner V, Pihlgren M, López-Deber M, Reis P, Hickman D, Adolfsson O, Chuard N, Ndao D, Borghgraef P, Devijver H, Van Leuven F, Pfeifer A, Muhs A. Efficacy and safety of a liposome-based vaccine against protein Tau, assessed in tau P301L mice that model tauopathy. PLoS One. 2013;8(8):e72301. doi: 10.1371/journal.pone.0072301. doi:10.1371/journal.pone.0072301.g005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toledano Furman N, Lupu-Haber Y, Bronshtein T, Kaneti L, Letko N, Weinstein E, Baruch L, Machluf M. Reconstructed stem cell nanoghosts: a natural tumor targeting platform. Nano Lett. 2013;13(7):3248–3255. doi: 10.1021/nl401376w. doi:10.1021/nl401376w. [DOI] [PubMed] [Google Scholar]
- Trajkovic K, Hsu C, Chiantia S, Rajendran L, Wenzel D, Wieland F, Schwille P, Brugger B, Simons M. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science. 2008;319(5867):1244–1247. doi: 10.1126/science.1153124. doi:10.1126/science.1153124. [DOI] [PubMed] [Google Scholar]
- Trumpfheller C, Longhi MP, Caskey M, Idoyaga J, Bozzacco L, Keler T, Schlesinger SJ, Steinman RM. Dendritic cell-targeted protein vaccines: a novel approach to induce T-cell immunity. J Intern Med. 2012;271(2):183–192. doi: 10.1111/j.1365-2796.2011.02496.x. doi:10.1111/j.1365-2796.2011.02496.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Umezu T, Ohyashiki K, Kuroda M, Ohyashiki JH. Leukemia cell to endothelial cell communication via exosomal miRNAs. Oncogene. 2013;32(22):2747–2755. doi: 10.1038/onc.2012.295. doi:10.1038/onc.2012.295. [DOI] [PubMed] [Google Scholar]
- Upadhyay K, Bhatt A, Mishra A, Dwarakanath B, Jain S, Schatz C, Le Meins J-F, Farooque A, Chandraiah G, Jain A, Misra A, Lecommandoux S. The intracellular drug delivery and anti tumor activity of doxorubicin loaded poly (gamma-benzyl l-glutamate)-b-hyaluronan polymersomes. Biomaterials. 2010;31(10):2882–2892. doi: 10.1016/j.biomaterials.2009.12.043. doi:10.1016/j.biomaterials.2009.12.043. [DOI] [PubMed] [Google Scholar]
- Valadi H, Ekström K, Bossios A, Sjöstrand M, Lee JJ, Lötvall JO. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol. 2007;9(6):654–659. doi: 10.1038/ncb1596. doi:10.1038/ncb1596. [DOI] [PubMed] [Google Scholar]
- Valencia K, Luis-Ravelo D, Bovy N, Anton I, Martinez-Canarias S, Zandueta C, Ormazabal C, Struman I, Tabruyn S, Rebmann V, De Las Rivas J, Guruceaga E, Bandres E, Lecanda F. miRNA cargo within exosome-like vesicle transfer influences metastatic bone colonization. Mol Oncol. 2014;8(3):689–703. doi: 10.1016/j.molonc.2014.01.012. doi:10.1016/j.molonc.2014.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Valenti R, Huber V, Filipazzi P, Pilla L, Sovena G, Villa A, Corbelli A, Fais S, Parmiani G, Rivoltini L. Human tumor-released microvesicles promote the differentiation of myeloid cells with transforming growth factor-beta-mediated suppressive activity on T lymphocytes. Cancer Res. 2006;66(18):9290–9298. doi: 10.1158/0008-5472.CAN-06-1819. doi:10.1158/0008-5472.can-06-1819. [DOI] [PubMed] [Google Scholar]
- van der Vlist EJ, Nolte-’t Hoen EN, Stoorvogel W, Arkesteijn GJ, Wauben MH. Fluorescent labeling of nano-sized vesicles released by cells and subsequent quantitative and qualitative analysis by high-resolution flow cytometry. Nat Protoc. 2012;7(7):1311–1326. doi: 10.1038/nprot.2012.065. doi:10.1038/nprot.2012.065. [DOI] [PubMed] [Google Scholar]
- van Dommelen S, Vader P, Lakhal S, Kooijmans SA, van Solinge W, Wood M, Schiffelers R. Microvesicles and exosomes: opportunities for cell-derived membrane vesicles in drug delivery. J Control Release. 2012;161(2):635–644. doi: 10.1016/j.jconrel.2011.11.021. doi:10.1016/j.jconrel.2011.11.021. [DOI] [PubMed] [Google Scholar]
- van Niel G, Porto-Carreiro I, Simoes S, Raposo G. Exosomes: a common pathway for a specialized function. J Biochem. 2006;140(1):13–21. doi: 10.1093/jb/mvj128. doi:10.1093/jb/mvj128. [DOI] [PubMed] [Google Scholar]
- Vekrellis K, Xilouri M, Emmanouilidou E, Rideout H, Stefanis L. Pathological roles of α-synuclein in neurological disorders. Lancet Neurol. 2011;10(11):1015–1025. doi: 10.1016/S1474-4422(11)70213-7. doi:10.1016/S1474-4422(11)70213-7. [DOI] [PubMed] [Google Scholar]
- Vella LJ, Sharples RA, Lawson VA, Masters CL, Cappai R, Hill AF. Packaging of prions into exosomes is associated with a novel pathway of PrP processing. J Pathol. 2007;211(5):582–590. doi: 10.1002/path.2145. doi:10.1002/path.2145. [DOI] [PubMed] [Google Scholar]
- Vella L, Sharples R, Nisbet R, Cappai R, Hill A. The role of exosomes in the processing of proteins associated with neurodegenerative diseases. Eur Biophys J. 2008;37(3):323–332. doi: 10.1007/s00249-007-0246-z. doi:10.1007/s00249-007-0246-z. [DOI] [PubMed] [Google Scholar]
- Verderio C, Muzio L, Turola E, Bergami A, Novellino L, Ruffini F, Riganti L, Corradini I, Francolini M, Garzetti L, Maiorino C, Servida F, Vercelli A, Rocca M, Dalla Libera D, Martinelli V, Comi G, Martino G, Matteoli M, Furlan R. Myeloid microvesicles are a marker and therapeutic target for neuroinflammation. Ann Neurol. 2012;72(4):610–624. doi: 10.1002/ana.23627. doi:10.1002/ana.23627. [DOI] [PubMed] [Google Scholar]
- Viaud S, Ploix S, Lapierre V, Thery C, Commere PH, Tramalloni D, Gorrichon K, Virault-Rocroy P, Tursz T, Lantz O, Zitvogel L, Chaput N. Updated technology to produce highly immunogenic dendritic cell-derived exosomes of clinical grade: a critical role of interferon-gamma. J Immunother. 2011;34(1):65–75. doi: 10.1097/CJI.0b013e3181fe535b. doi:10.1097/CJI.0b013e3181fe535b. [DOI] [PubMed] [Google Scholar]
- Villarroya-Beltri C, Gutiérrez-Vázquez C. Sumoylated hnRNPA2B1 controls the sorting of miRNAs into exosomes through binding to specific motifs. Nat Commun. 2013;4:2980. doi: 10.1038/ncomms3980. doi:10.1038/ncomms3980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vlassov A, Magdaleno S, Setterquist R, Conrad R. Exosomes: current knowledge of their composition, biological functions, and diagnostic and therapeutic potentials. Biochim Biophys Acta. 2012;7:940–948. doi: 10.1016/j.bbagen.2012.03.017. doi:10.1016/j.bbagen.2012.03.017. [DOI] [PubMed] [Google Scholar]
- Von Bartheld CS, Altick AL. Multivesicular bodies in neurons: distribution, protein content, and trafficking functions. Prog Neurobiol. 2011;93(3):313–340. doi: 10.1016/j.pneurobio.2011.01.003. doi:10.1016/j.pneurobio.2011.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wahlgren J, De LKT, Brisslert M, Vaziri Sani F, Telemo E, Sunnerhagen P, Valadi H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012;40(17):e130. doi: 10.1093/nar/gks463. doi:10.1093/nar/gks463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walsh A, Yin H, Erben C, Wood M, Turberfield A. DNA cage delivery to mammalian cells. ACS Nano. 2011;5(7):5427–5432. doi: 10.1021/nn2005574. doi:10.1021/nn2005574. [DOI] [PubMed] [Google Scholar]
- Wang H, Li F, Du C, Wang H, Mahato R, Huang Y. Doxorubicin and lapatinib combination nanomedicine for treating resistant breast cancer. Mol Pharma. 2014a doi: 10.1021/mp400687w. in press. doi:10.1021/mp400687w. [DOI] [PubMed] [Google Scholar]
- Wang M, Zhao C, Shi H, Zhang B, Zhang L, Zhang X, Wang S, Wu X, Yang T, Huang F, Cai J, Zhu Q, Zhu W, Qian H, Xu W. Deregulated microRNAs in gastric cancer tissue-derived mesenchymal stem cells: novel biomarkers and a mechanism for gastric cancer. Br J Cancer. 2014b;110(5):1199–1210. doi: 10.1038/bjc.2014.14. doi:10.1038/bjc.2014.14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang-Johanning F, Frost AR, Jian B, Epp L, Lu DW, Johanning GL. Quantitation of HERV-K env gene expression and splicing in human breast cancer. Oncogene. 2003;22(10):1528–1535. doi: 10.1038/sj.onc.1206241. doi:10.1038/sj.onc.1206241. [DOI] [PubMed] [Google Scholar]
- Wasungu L, Hoekstra D. Cationic lipids, lipoplexes and intracellular delivery of genes. J Control Release. 2006;116(2):255–264. doi: 10.1016/j.jconrel.2006.06.024. doi:10.1016/j.jconrel.2006.06.024. [DOI] [PubMed] [Google Scholar]
- Waterhouse DN, Tardi PG, Mayer LD, Bally MB. A comparison of liposomal formulations of doxorubicin with drug administered in free form: changing toxicity profiles. Drug Saf. 2001;24(12):903–920. doi: 10.2165/00002018-200124120-00004. doi:10.0000/095372800232108. [DOI] [PubMed] [Google Scholar]
- Wen Z, Yan Z, Hu K, Pang Z, Cheng X, Guo L, Zhang Q, Jiang X, Fang L, Lai R. Odorranalectin-conjugated nanoparticles: preparation, brain delivery and pharmaco-dynamic study on Parkinson’s disease following intranasal administration. J Control Release. 2011;151(2):131–138. doi: 10.1016/j.jconrel.2011.02.022. doi:10.1016/j.jconrel.2011.02.022. [DOI] [PubMed] [Google Scholar]
- Wiley RD, Gummuluru S. Immature dendritic cell-derived exosomes can mediate HIV-1 trans infection. Proc Natl Acad Sci USA. 2006;103(3):738–743. doi: 10.1073/pnas.0507995103. doi:10.1073/pnas.0507995103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wittmann J, Jäck H-M. Serum microRNAs as powerful cancer biomarkers. Biochim Biophys Acta. 2010;1806(2):200–207. doi: 10.1016/j.bbcan.2010.07.002. doi:10.1016/j.bbcan.2010.07.002. [DOI] [PubMed] [Google Scholar]
- Wu J, Huang W, He Z. Dendrimers as carriers for siRNA delivery and gene silencing: a review. Sci World J. 2013;2013:630654. doi: 10.1155/2013/630654. doi:10.1155/2013/630654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wubbolts R, Leckie RS, Veenhuizen PT, Schwarzmann G, Mobius W, Hoernschemeyer J, Slot JW, Geuze HJ, Stoorvogel W. Proteomic and biochemical analyses of human B cell-derived exosomes. Potential implications for their function and multivesicular body formation. J Biol Chem. 2003;278(13):10963–10972. doi: 10.1074/jbc.M207550200. doi:10.1074/jbc.M207550200. [DOI] [PubMed] [Google Scholar]
- Xia C-F, Boado R, Zhang Y, Chu C, Pardridge W. Intravenous glial-derived neurotrophic factor gene therapy of experimental Parkinson’s disease with Trojan horse liposomes and a tyrosine hydroxylase promoter. J Gene Med. 2008;10(3):306–315. doi: 10.1002/jgm.1152. doi:10.1002/jgm.1152. [DOI] [PubMed] [Google Scholar]
- Xiao D, Ohlendorf J, Chen Y, Taylor DD, Rai SN, Waigel S, Zacharias W, Hao H, McMasters KM. Identifying mRNA, microRNA and protein profiles of melanoma exosomes. PLoS One. 2012;7(10):e46874. doi: 10.1371/journal.pone.0046874. doi:10.1371/journal.pone.0046874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xin H, Li Y, Buller B, Katakowski M, Zhang Y, Wang X, Shang X, Zhang ZG, Chopp M. Exosome-mediated transfer of miR-133b from multipotent mesenchymal stromal cells to neural cells contributes to neurite outgrowth. Stem Cells. 2012;30(7):1556–1564. doi: 10.1002/stem.1129. doi:10.1002/stem.1129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xin H, Li Y, Cui Y, Yang JJ, Zhang ZG, Chopp M. Systemic administration of exosomes released from mesenchymal stromal cells promote functional recovery and neurovascular plasticity after stroke in rats. J Cereb Blood Flow Metab. 2013a;33(11):1711–1715. doi: 10.1038/jcbfm.2013.152. doi:10.1038/jcbfm.2013.152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xin H, Li Y, Liu Z, Wang X, Shang X, Cui Y, Gang Zhang Z, Chopp M. Mir-133b promotes neural plasticity and functional recovery after treatment of stroke with multi-potent mesenchymal stromal cells in rats via transfer of exosome-enriched extracellular particles. Stem Cells. 2013b doi: 10.1002/stem.1409. doi:10.1002/stem.1409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu W, Ling P, Zhang T. Polymeric micelles, a promising drug delivery system to enhance bioavailability of poorly water-soluble drugs. J Drug Deliv. 2013;2013:340315. doi: 10.1155/2013/340315. doi:10.1155/2013/340315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamada N, Nakagawa Y, Tsujimura N, Kumazaki M, Noguchi S, Mori T, Hirata I, Maruo K, Akao Y. Role of intracellular and extracellular microRNA-92a in colorectal cancer. Transl Oncol. 2013;6(4):482–492. doi: 10.1593/tlo.13280. doi:10.1593/tlo.13280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang M, Chen J, Su F, Yu B, Su F, Lin L, Liu Y, Huang J-D, Song E. Microvesicles secreted by macrophages shuttle invasion-potentiating microRNAs into breast cancer cells. Mol cancer. 2011;10:117. doi: 10.1186/1476-4598-10-117. doi:10.1186/1476-4598-10-117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu X, Harris S, Levine A. The regulation of exosome secretion: a novel function of the p53 protein. Cancer Res. 2006;66(9):4795–4801. doi: 10.1158/0008-5472.CAN-05-4579. doi:10.1158/0008-5472.CAN-05-4579. [DOI] [PubMed] [Google Scholar]
- Yu X, Riley T, Levine A. The regulation of the endosomal compartment by p53 the tumor suppressor gene. FEBS J. 2009;276(8):2201–2212. doi: 10.1111/j.1742-4658.2009.06949.x. doi:10.1111/j.1742-4658.2009.06949.x. [DOI] [PubMed] [Google Scholar]
- Yu Y, Pang Z, Lu W, Yin Q, Gao H, Jiang X. Self-assembled polymersomes conjugated with lactoferrin as novel drug carrier for brain delivery. Pharmaceut Res. 2012;29(1):83–96. doi: 10.1007/s11095-011-0513-7. doi:10.1007/s11095-011-0513-7. [DOI] [PubMed] [Google Scholar]
- Yu L, Yang F, Jiang L, Chen Y, Wang K, Xu F, Wei Y, Cao X, Wang J, Cai Z. Exosomes with membrane-associated TGF-beta1 from gene-modified dendritic cells inhibit murine EAE independently of MHC restriction. Eur J Immunol. 2013;43(9):2461–2472. doi: 10.1002/eji.201243295. doi:10.1002/eji.201243295. [DOI] [PubMed] [Google Scholar]
- Yuyama K, Yamamoto N, Yanagisawa K. Accelerated release of exosome-associated GM1 ganglioside (GM1) by endocytic pathway abnormality: another putative pathway for GM1-induced amyloid fibril formation. J Neurochem. 2008;105(1):217–224. doi: 10.1111/j.1471-4159.2007.05128.x. doi:10.1111/j.1471-4159.2007.05128.x. [DOI] [PubMed] [Google Scholar]
- Yuyama K, Sun H, Mitsutake S, Igarashi Y. Sphingolipid-modulated exosome secretion promotes clearance of amyloid-β by microglia. J Biol Chem. 2012;287(14):10977–10989. doi: 10.1074/jbc.M111.324616. doi:10.1074/jbc.M111.324616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zeelenberg IS, Ostrowski M, Krumeich S, Bobrie A, Jancic C, Boissonnas A, Delcayre A, Le Pecq JB, Combadière B, Amigorena S, Théry C. Targeting tumor antigens to secreted membrane vesicles in vivo induces efficient antitumor immune responses. Cancer Res. 2008;68(4):1228–1235. doi: 10.1158/0008-5472.CAN-07-3163. doi:10.1158/0008-5472.can-07-3163. [DOI] [PubMed] [Google Scholar]
- Zhan R, Leng X, Liu X, Wang X, Gong J, Yan L. Heat shock protein 70 is secreted from endothelial cells by a non-classical pathway involving exosomes. Biochem Biophys Res Commun. 2009;387(2):229–233. doi: 10.1016/j.bbrc.2009.06.095. doi:10.1016/j.bbrc.2009.06.095. [DOI] [PubMed] [Google Scholar]
- Zhan C, Wei X, Qian J, Feng L, Zhu J, Lu W. Co-delivery of TRAIL gene enhances the anti-glioblastoma effect of paclitaxel in vitro and in vivo. J Control Release. 2012;160(3):630–636. doi: 10.1016/j.jconrel.2012.02.022. doi:10.1016/j.jconrel.2012.02.022. [DOI] [PubMed] [Google Scholar]
- Zhang Y, Zhang Y, Bryant J, Charles A, Boado R. Intravenous RNA interference gene therapy targeting the human epidermal growth factor receptor prolongs survival in intracranial brain cancer. Clin Cancer Res. 2004;10(11):3667–3677. doi: 10.1158/1078-0432.CCR-03-0740. doi:10.1158/1078-0432.CCR-03-0740. [DOI] [PubMed] [Google Scholar]
- Zhang Y, Liu D, Chen X, Li J, Li L, Bian Z, Sun F, Lu J, Yin Y, Cai X, Sun Q, Wang K, Ba Y, Wang Q, Wang D, Yang J, Liu P, Xu T, Yan Q, Zhang J, Zen K, Zhang C-Y. Secreted monocytic miR-150 enhances targeted endothelial cell migration. Mol Cell. 2010;39(1):133–144. doi: 10.1016/j.molcel.2010.06.010. doi:10.1016/j.molcel.2010.06.010. [DOI] [PubMed] [Google Scholar]
- Zhang X-X, McIntosh T, Grinstaff M. Functional lipids and lipoplexes for improved gene delivery. Biochimie. 2012;94(1):42–58. doi: 10.1016/j.biochi.2011.05.005. doi:10.1016/j.biochi.2011.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu J, Shi X. Dendrimer-based nanodevices for targeted drug delivery applications. J Mater Chem B. 2013;1(34):4199. doi: 10.1039/c3tb20724b. doi:10.1039/c3tb20724b. [DOI] [PubMed] [Google Scholar]
- Zhu G, Hu R, Zhao Z, Chen Z, Zhang X, Tan W. Non-canonical self-assembly of multifunctional DNA nanoflowers for biomedical applications. J Am Chem Soc. 2013;135(44):16438–16445. doi: 10.1021/ja406115e. doi:10.1021/ja406115e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhuang X, Xiang X, Grizzle W, Sun D, Zhang S, Axtell RC, Ju S, Mu J, Zhang L, Steinman L, Miller D, Zhang HG. Treatment of brain inflammatory diseases by delivering exosome encapsulated anti-inflammatory drugs from the nasal region to the brain. Mol Ther. 2011;19(10):1769–1779. doi: 10.1038/mt.2011.164. doi:10.1038/mt.2011.164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zia ur R, Dick H, Inge SZ. Mechanism of polyplex- and lipoplex-mediated delivery of nucleic acids: real-time visualization of transient membrane destabilization without endosomal lysis. ACS Nano. 2013;7(5):3767–3777. doi: 10.1021/nn3049494. doi:10.1021/nn3049494. [DOI] [PubMed] [Google Scholar]
- Zomer A, Vendrig T, Hopmans ES, van Eijndhoven M, Middeldorp JM, Pegtel DM. Exosomes: fit to deliver small RNA. Commun Integr Biol. 2010;3(5):447–450. doi: 10.4161/cib.3.5.12339. doi:10.4161/cib.3.5.12339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zwaal RF, Schroit AJ. Pathophysiologic implications of membrane phospholipid asymmetry in blood cells. Blood. 1997;89(4):1121–1132. [PubMed] [Google Scholar]